Anda di halaman 1dari 171

Calculation Tools for Estimating

Greenhouse Gas Emissions


from Pulp and Paper Mills

Version 1.0

A Project of
The Climate Change Working Group
of
The International Council of Forest and Paper Associations (ICFPA)
with special contributions from
Paper Manufacturers Association of South Africa
Japan Paper Association
Forest Products Association of Canada
Confederation of European Paper Industries
Chilean Forest and Industry Association
Australian Paper Industry Council
American Forest and Paper Association

Prepared by
NCASI, Research Triangle Park, NC, USA

Questions or comments on this material can be directed to


Reid Miner
NCASI
P.O. Box 13318
Research Triangle Park, NC, USA 27709-3318
phone + 919-558-1991
fax +919-558-1998
e-mail rminer@ncasi.org
CALCULATION TOOLS FOR ESTIMATING GREENHOUSE GAS EMISSIONS
FROM PULP AND PAPER MILLS

EXECUTIVE SUMMARY

In early 2001, a group of pulp and paper industry associations from around the globe (the
International Council of Forest and Paper Associations) decided to jointly support a project
to develop international tools for estimating greenhouse gas (GHG) emissions from pulp and
paper mills. The group retained the National Council for Air and Stream Improvement, Inc.
(NCASI) to review existing greenhouse gas inventory protocols and to develop calculation
tools to assist companies in preparing GHG inventories. The results of the review and the
calculation tools are contained in this report. It is intended that these industry-specific tools
be used in conjunction with an accepted GHG accounting protocol such as the “Greenhouse
Gas Protocol” issued by the World Resources Institute/World Business Council for
Sustainable Development (WRI/WBCSD) or other accepted protocols for corporate GHG
inventories.

These tools reflect many of the features of well-known and widely accepted protocols. In
addition, they anticipate a number of questions that pulp and paper mills must address when
preparing facility-level or company-level inventories. A special effort has been made to
ensure that the tools are consistent with protocols issued by the Intergovernmental Panel on
Climate Change (IPCC) and the WRI/WBCSD.

These tools estimate CO2 emissions from fossil fuel combustion based on the carbon content
of the fuel (or a comparable emission factor) and the amount burned. Carbon dioxide
emissions from biomass combustion are not counted as greenhouse gas emissions, a
convention common to most of the protocols examined in this review, but can be reported as
“supporting information” as called for in the WRI/WBCSD GHG Protocol. Methane and
nitrous oxide emissions from combustion processes, both fossil fuel and biomass, are
estimated using emission factors. Methods are presented for estimating the fossil-CO2,
methane, and nitrous oxide emissions from lime kilns and calciners. Greenhouse gas
emissions from mill landfills and wastewater treatment plants are estimated using methods
derived from those suggested by IPCC, as are emissions from vehicles and other fossil fuel-
fired equipment. In all cases, however, companies may use site-specific information where it
yields more accurate estimates of GHG emissions than the tools outlined in this report.

Using these tools, indirect emissions related to imports of electricity or steam and emissions
attributable to exports of electricity or steam are included in the inventory results but tracked
separately from direct emissions. The emissions from combined heat and power (CHP) plants
are allocated using the WRI/WBCSD “efficiency method.”

The calculation tools envision that companies will often construct inventory boundaries that
include emissions from (a) all on-site “core” pulp and paper making operations, regardless of
the ownership of the emission source; and (b) other company-owned sources included in
widely accepted GHG inventory protocols (e.g., company-owned truck fleets). It is
understood, however, that companies will select inventory boundaries suited to the objectives
of the inventory. For most mills, the GHG profile will be dominated by stationary fossil fuel
combustion emissions and emissions attributable to purchases of power and steam, emissions
that are discussed in Sections 8 and 12.

To aid in interpreting the results of the inventory, these tools recommend that the inventory
results include a list of the operations contained within the boundaries of the inventory and a
list of emission factors used to estimate emissions. The format suggested for presenting the
results of the inventory allows a company to separate the emissions that are “owned” by the
company (direct emissions) from those that are not (indirect emissions). The company is free
to select the method for determining the ownership of emissions, but the method must be
explained in the inventory results. The user is directed to the WRI/WBCSD GHG Protocol
for guidance on how to determine ownership of emissions from partly-owned or partly-
controlled sources.

An Excelâ workbook is available to assist in performing the calculations described in this


report.
Acknowledgements

This document (and the Excelâ file that accompanies it) have undergone a peer review
which was organized by WRI/WBCSD. NCASI and ICFPA acknowledge the contributions
of the peer review participants:

· Adam Costanza, International Paper Company


· Chris Laughran and Terry Shires, URS Corporation, on behalf of Karen Ritter
and the American Petroleum Institute GHG Emissions Methodology Workgroup
· Inna Gritsevich, Center for Energy Efficiency (CENEf) in Russia

The assistance from members of the ICFPA Climate Change Working Group and Technical
Experts Working Group are also gratefully acknowledged.
CONTENTS

1.0 INTRODUCTION ............................................................................................................1

2.0 FOREST PRODUCTS INDUSTRY GREENHOUSE GAS EMISSIONS IN


PERSPECTIVE.................................................................................................................2

3.0 RELATIONSHIP OF THESE CALCULATION TOOLS TO OTHER


GREENHOUSE GAS PROTOCOLS...............................................................................2

4.0 OVERVIEW OF THE CALCULATION TOOLS ...........................................................4

4.1 Purpose of the Calculation Tools.............................................................................4

4.2 Steps Involved in Applying the Calculation Tools ..................................................4

4.3 Data Quality.............................................................................................................6

4.4 Units.........................................................................................................................7

5.0 IDENTIFYING CORE PULP AND PAPER OPERATIONS TO BE INCLUDED


WITHIN THE OPERATIONAL BOUNDARIES OF THE INVENTORY ....................8

6.0 DETERMINING THE ORGANIZATIONAL BOUNDARIES OF THE


INVENTORY AND OWNERSHIP OF EMISSIONS ...................................................10

7.0 MATERIALITY AND INSIGNIFICANT EMISSIONS ...............................................11

8.0 GREENHOUSE GAS EMISSIONS FROM STATIONARY FOSSIL FUEL


COMBUSTION ..............................................................................................................13

8.1 Carbon Dioxide......................................................................................................13

8.2 Methane and Nitrous Oxide ...................................................................................14

9.0 EMISSIONS FROM KRAFT MILL LIME KILNS AND CALCINERS......................19

10.0 CARBON DIOXIDE EMISSIONS FROM MAKE-UP CHEMICALS ........................22

11.0 EMISSIONS FROM STATIONARY COMBUSTION OF BIOMASS FUELS ...........23

11.1 Climate-Neutral Carbon Dioxide Emissions from Burning Biomass Fuels ..........23

11.2 Methane and Nitrous Oxide Emissions from Burning Biomass Fuels ..................24

12.0 EMISSIONS ATTRIBUTABLE TO IMPORTS AND EXPORTS OF


ELECTRICITY AND STEAM.......................................................................................26

12.1 Emission Factors for Purchased Power and Steam................................................26


12.2 Netting Imports and Exports..................................................................................27

12.3 Electricity Imports .................................................................................................27

12.4 Electricity Exports .................................................................................................28

12.5 Steam Imports ........................................................................................................29

12.6 Steam and Hot Water Exports................................................................................29

12.7 Allocating Emissions from Combined Heat and Power (CHP) Systems ..............29

13.0 GREENHOUSE GAS EMISSIONS FROM VEHICLES AND


MISCELLANEOUS FOSSIL FUEL-FIRED EQUIPMENT.........................................33

13.1 Greenhouse Gases from On-Road Vehicles ..........................................................33

13.2 Greenhouse Gases from Off-Road Vehicles and Equipment ................................33

14.0 GREENHOUSE GAS EMISSIONS FROM MILL WASTE IN LANDFILLS.............36

14.1 Using Data from Landfill Gas Collection Systems................................................37

14.2 Estimating Landfill Methane Emissions at Mill Landfills without Gas


Collection Data ......................................................................................................38

15.0 GREENHOUSE GAS EMISSIONS FROM ANAEROBIC TREATMENT OF


WASTEWATER OR SLUDGE .....................................................................................42

15.1 Anaerobic Treatment Operations where Off-Gases are Captured .........................42

15.2 Anaerobic Treatment Operations where Off-Gases are Released to the


Atmosphere ............................................................................................................42

16.0 PRESENTING THE RESULTS OF THE INVENTORY..............................................43

REFERENCES ........................................................................................................................54
ANNEXES

A GREENHOUSE GAS EMISSIONS FROM STATIONARY FOSSIL FUEL


COMBUSTION – OVERVIEW OF METHODS IN EXISTING PROTOCOLS ........A1

B GREENHOUSE GAS EMISSIONS FROM KRAFT MILL LIME KILNS AND


CALCINERS .................................................................................................................B1

C CH4 AND N2O EMISSIONS FROM BIOMASS COMBUSTION – OVERVIEW


OF METHODS IN EXISTING PROTOCOLS .............................................................C1

D GREENHOUSE GAS EMISSIONS ATTRIBUTABLE TO IMPORTS AND


EXPORTS OF POWER AND STEAM – OVERVIEW OF METHODS IN
EXISTING PROTOCOLS.............................................................................................D1

E ALLOCATING GREENHOUSE GAS EMISSIONS FROM COMBINED HEAT


AND POWER (CHP) SYSTEMS – RECOMMENDED GUIDANCE AND
REVIEW OF METHODS.............................................................................................. E1

F GREENHOUSE GASES FROM VEHICULAR TRAFFIC AND MACHINERY


– OVERVIEW OF METHODS IN EXISTING PROTOCOLS.................................... F1

G GREENHOUSE GASES FROM WASTE MANAGEMENT AT PULP AND


PAPER MILLS – RECOMMENDED APPROACH AND REVIEW OF
EXISTING METHODS.................................................................................................G1

H ALLOCATING GREENHOUSE GAS EMISSIONS FROM PARTLY OWNED


OR PARTLY CONTROLLED SOURCES – OVERVIEW OF METHODS IN
EXISTING PROTOCOLS.............................................................................................H1

I SUPPORTING INFORMATION ON BIOMASS ......................................................... I1

J REFERENCES FOR ANNEXES A THROUGH I ........................................................ J1


TABLES

Table 1. Emission Factor Ranges Useful in Identifying Significant and Insignificant


Sources of GHGs ...................................................................................................12

Table 2. IPCC Default CO2 Emission Factors for Fossil Fuels ...........................................14

Table 3. IPCC Tier 2 Uncontrolled CH4 and N2O Emission Factors for Industrial
Boilers ....................................................................................................................16

Table 4. Emission Factors for Kraft Mill Lime Kilns and Calciners...................................20

Table 5 Emission Factors for Calcium- and Sodium-Carbonate Make-up in the Pulp
Mill.........................................................................................................................22

Table 6. Emission Factors for CH4 and N2O from Biomass Combustion ...........................25

Table 7. Default Fuel Economy and Emission Factors for Different Types of Mobile
Sources and Activity Data .....................................................................................34

Table 8. Fuel Consumption-Based Emission Factors for Non-Road Mobile Sources


and Machinery .......................................................................................................35

Table 9. Recommended Default Values for k and L0 for Estimating Mill Landfill
Methane Emissions ................................................................................................39

Table 10. Table to Report Operational Boundaries of the Inventory.....................................45

Table 11. Table to Report GHG Inventory Results – Direct Emissions ................................46

Table 12. Table to Report GHG Inventory Results – Indirect Emissions..............................47

Table 13. Table to Report Emission Factors Used to Prepare the Inventory.........................48

Table 14. Example GHG Inventory Results – Operational Boundaries of the Inventory......49

Table 15. Example GHG Inventory Results – Direct Emissions...........................................51

Table 16. Example GHG Inventory Results – Indirect Emissions ........................................52

Table 17. Example GHG Inventory Results – Emission Factors Used to Prepare the
Inventory................................................................................................................53

FIGURES

Figure 1. Schematic for Example Inventory Results.............................................................50


CALCULATION TOOLS FOR ESTIMATING GREENHOUSE GAS EMISSIONS
FROM PULP AND PAPER MILLS

1.0 INTRODUCTION

Responding to the need for improved methods for estimating greenhouse gas (GHG)
emissions from pulp and paper mills, in 2001 the International Council of Forest and Paper
Associations (ICFPA) agreed to develop international tools to
· enable harmonized collection of credible, transparent, and comparable data worldwide
· address the forest product industry’s unique attributes
· establish a framework that would assist in implementing a variety of programs that might
make use of carbon inventory data

To accomplish this, the ICFPA Climate Change Working Group retained the research
institute National Council for Air and Stream Improvement, Inc. (NCASI) to review existing
GHG protocols and assist the industry in developing calculation tools for estimating GHG
emissions.

The results of that effort are contained in this report. The calculation tools are described in
the body of the report. The Annexes summarize relevant features of the calculation methods
used in a number of existing GHG protocols and provide additional detail on estimation
methods. The Annexes contain many of the country-specific data and factors published by
national authorities. The information in the Annexes was current as of the end of 2001.

This material encompasses only manufacturing-related emissions from pulp and paper
production. Issues related to carbon sinks or forest sequestration are not addressed.1

These calculation tools will assist companies in preparing GHG emission inventories for a
number of purposes including internal company benchmarking, public reporting, product
profiles, and carbon trading. The rules governing the development of a GHG inventory,
however, can vary substantially from one program to another, so the user of these tools
should always be familiar with the requirements imposed by the intended use of the
inventory results.

These industry-specific tools should be used in conjunction with an accepted GHG


accounting protocol such as the “Greenhouse Gas Protocol” issued by the World Resources
Institute/World Business Council for Sustainable Development (WRI/WBCSD) or other
accepted protocols for corporate GHG inventories. These protocols provide valuable
information on issues ranging from defining the objectives for a GHG inventory to options
for verifying the results – issues beyond the scope of the industry-specific calculation tools in
this report.

1
IPCC has developed methods that countries are using to characterize sequestration (IPCC 1997a, b, c, 2000b),
and a great deal of work is underway to improve the understanding of sequestration and its measurement.
Some of these studies are summarized by Skog and Nicholson 1998; Apps et al. 1999; Matthews 1996; and
Birdsey 1996.

1
2.0 FOREST PRODUCTS INDUSTRY GREENHOUSE GAS EMISSIONS IN
PERSPECTIVE

The forest products industry has an important and complex role in the global carbon cycle.
Forests supply the industry’s primary raw material. The sustainable management of these
forests sequesters massive amounts of carbon and provides needed products that contribute to
significant carbon pools during their use and after being discarded. In addition, forests
provide multiple environmental, social, and economic benefits.

Efforts to expand the amount of forested land are increasing carbon storage in most of the
developed world and new plantations are being established in many areas of the developing
world. Research is ongoing to identify forest management practices capable of optimizing
carbon storage in existing forests while maintaining or enhancing forest productivity and
protecting the environment.

Carbon is also stored when forests are managed to produce needed products because many of
these products store carbon for extended periods of time as they are used and after disposal.
Recycling is an important part of the carbon cycle since it can help extend the time during
which carbon is stored in products. It has been estimated that, on a global basis, the amount
of carbon stored in forest products is increasing by 139 million metric tons of carbon per year
(Winjum et al. 1998).

The forest products industry relies heavily on climate-neutral biomass fuels that reduce
atmospheric levels of greenhouse gases by displacing fossil fuels. In a number of countries,
more than half of the industry’s energy requirements are met using biomass fuels. Forest
products that cannot be economically recycled provide another source of biomass fuels.

The pulp and paper industry is one of the global leaders in the use of combined heat and
power (CHP) systems, also called cogeneration systems. CHP systems produce electrical
power and thermal energy from the same fuel, yielding twice as much or more usable energy
from the fuel than normal methods for generating power and steam. This reduces GHG
emissions by reducing the demand for fossil fuels. The pulp and paper industries in some
countries derive more than half of their energy from CHP systems.

The industry’s interactions with the global carbon cycle are extensive and complex. It is
important, therefore, that the industry’s GHG emissions not be viewed in isolation. It is only
within the context of the overall forest products carbon cycle that the significance of the
industry’s emissions can be properly evaluated.

3.0 RELATIONSHIP OF THESE CALCULATION TOOLS TO OTHER


GREENHOUSE GAS PROTOCOLS

There are many protocols for estimating and reporting GHG emissions. Most of the existing
protocols are based on a common set of general principles with differences primarily
attributable to the differing purposes of the protocols (e.g., national inventories, corporate
inventories, project accounting, etc.). The general principles for GHG inventory
development are important and should be addressed in preparing any inventory of GHG

2
emissions. This report, however, devotes relatively little attention to such issues because
these principles are generic and information is available in a variety of other places.

Some especially helpful sources of general information on inventory preparation are:


· the Intergovernmental Panel on Climate Change (IPCC) (IPCC 1997a, b, c, 2000a)
· the World Resources Institute/World Business Council for Sustainable Development
(WRI/WBCSD) (WRI 2001a)
· the PEW Center on Global Climate Change (Loreti, Wescott, and Isenberg 2000; Loreti,
Foster, and Obbagy 2001)

The WRI/WBCSD and PEW Center documents are especially relevant for companies using
these calculation tools because they focus on company-level reporting. Instructions on how
to obtain these documents are included in the literature citations. The WRI/WBCSD and
PEW Center documents provide helpful information on the following generic, but important,
issues:
· GHG accounting and reporting principles (e.g., organizational and operational boundary
determination, relevance, completeness, consistency, transparency, and accuracy
· defining corporate objectives for inventories (e.g., public reporting, voluntary initiatives,
carbon trading)
· establishing baselines and accounting for reductions
· managing inventory quality
· verification

The pulp and paper industry calculation tools in this report are intended to assist companies
in developing data that can be used to fulfill the requirements of a number of protocols,
including those derived from IPCC guidelines and best practices and the WRI/WBCSD GHG
Protocol.

Given the widespread acceptance of the WRI/WBCSD GHG Protocol, it is important to note
that there is one area where these calculation tools may yield information that is not
completely adequate for reporting under the WRI/WBCSD GHG Protocol. In specific, the
GHG Protocol suggests that companies report emissions of hydrofluorocarbons (HFCs) from
air conditioning and refrigeration, but HFC emissions are not addressed in these pulp and
paper mill calculation tools. WRI/WBCSD has not yet published calculation tools for these
emissions, but IPCC has issued guidance that companies wanting to estimate these emissions
might find helpful (IPCC 1997c, Section 2.17.4.2).

There are other differences between the WRI/WBCSD GHG Protocol and the calculation
tools presented here, but the differences result in these tools providing additional information
not required by the GHG Protocol or providing it in a slightly different format.

Perhaps the only significant variance from IPCC’s recommended approaches is in the area of
landfill emissions. IPCC’s approach relies on generic estimation methods, while these tools
suggest that site-specific landfill gas collection data can often be used as the basis for the
estimates where these data are available.

3
4.0 OVERVIEW OF THE CALCULATION TOOLS

4.1 Purpose of the Calculation Tools

These tools have been developed to help achieve the following outcomes:
· ensure credible and transparent emissions estimates for GHGs of potential significance to
pulp and paper mills; i.e., fossil-CO2, CH4 (methane), and N2O (nitrous oxide)
· provide companies with methods that are tailored to the conditions likely to be
encountered at pulp and paper mills and are easily implemented
· minimize the potential for inappropriate mill-to-mill, company-to-company, and
international comparisons of GHG emissions
· facilitate the tracking of improvements at individual sites or companies
· facilitate carbon trading
· be as consistent as possible with recognized international and national protocols

Although the tools are focused on pulp and paper mills, they are generally applicable to wood
products facilities as well. These tools do not deal with operations unique to wood products
manufacturing, but NCASI will be developing a set of tools tailored to the needs of those
facilities.

4.2 Steps Involved in Applying the Calculation Tools

In general terms, the calculation tools involve the user’s performing the following steps, most
of which are described in detail in the following sections of this report.

4.2.1 Determining the Objectives of the Inventory

The WRI/WBCSD GHG Protocol can help companies understand the variety of uses for
GHG inventory results (WRI 2001a). The way the inventory is designed and conducted will
depend largely on the intended uses of the results. Before undertaking a GHG inventory,
therefore, companies should assure themselves that the methods used to develop the
inventory meet the requirements imposed by the objectives of the inventory.

4.2.2 Identifying Boundary Conditions

There are two types of boundaries that must be considered in a GHG inventory – operational
boundaries and organizational boundaries. The operational boundaries define the operations
and emission sources that need to be included to satisfy the objectives of the inventory. The
organizational boundaries dictate the ownership of the emissions and determine whether the
emissions are “direct” or “indirect.”

Direct and indirect emissions are defined as follows (WRI 2001a):


· Direct emissions are “emissions from sources that are owned or controlled by the
reporting company.”
· Indirect emissions are “emissions that are a consequence of the activities of the reporting
company but occur from sources owned or controlled by another company.”

4
Of course, there is an almost endless chain of upstream and downstream “consequences”
that, at least theoretically, can be connected to a company’s activities. GHG protocols,
however, usually address only a limited set of indirect emissions – those associated with
transfers of electrical power, steam, and heat – and these are the indirect emissions addressed
in these calculation tools.

The tools have been developed anticipating that most companies will select operational
boundaries that include:
· direct emissions from on-site core operations (e.g., company-owned power boilers)
· direct emissions from off-site core operations (e.g., company-owned harvesting
equipment)
· indirect emissions related to imports of power or steam (including those from outsourced
power islands)
· emissions attributable to exported power or steam
· indirect emissions from on-site core operations not involving power and steam transfers
(e.g., outsourced but on-site wastewater treatment operations)

Core operations are those associated with the pulp and papermaking process. Emissions
from operations that are not part of the normal pulp and papermaking process are not
included within the scope of these tools, although companies may sometimes need to include
them to satisfy the objectives of the inventory.

4.2.3 Estimating Emissions from Sources within the Operational Boundaries

The next step is estimating the GHG emissions. The calculation tools in this report address:
· CO2 emissions from stationary fossil fuel combustion
· CH4 and N2O emissions from fossil fuel-fired units, recovery furnaces, biomass boilers,
and lime kilns
· CO2 emissions from make-up CaCO3 or Na2CO3 used in the pulp mill
· CO2 emissions from on-site vehicles and machinery
· CH4 emissions attributable to mill wastes in landfills and anaerobic waste treatment
operations
· emissions from other off-site core operations (e.g., company-owned harvesting
equipment and company-owned truck fleets)
· fossil fuel-derived CO2 exported to satellite precipitated calcium carbonate (PCC) plants
· imports of CO2 (e.g., for pH neutralization)
· GHG emissions associated with power and steam imports
· GHG emissions attributable to power and steam exports

Emissions of climate-neutral biomass-derived CO2 are addressed in Annex I.

Although not addressed in these calculation tools, fuel gas system piping may have methane
equipment leaks (e.g., piping associated with a natural gas boiler). Where users wish to
address such fugitive emissions, which would most likely be small in relation to the GHG
emissions categories listed above, more information can be obtained in the EPA publication
Protocol for Equipment Leak Emission Estimates (USEPA 1995).

5
4.2.4 Determining Organizational Boundaries and Ownership of Emissions

In addition to setting operational boundaries, the company will need to identify the
organizational boundaries, or in other words, the boundaries that dictate ownership of
emissions from the sources within the operational boundaries. A company’s organizational
boundaries will include those on-site and off-site emissions from sources that the company
“owns or controls” (determined by using one of a variety of methods discussed later).
Indirect emissions are outside the organizational boundaries of the inventory because, by
definition, indirect emissions “occur from sources owned or controlled by another company”
(WRI 2001a). Additional discussion of operational and organizational boundaries is
presented below and is available in the WRI/WBCSD GHG Protocol (WRI 2001a).

4.2.5 Presenting Results

The calculation tools emphasize the disaggregated and transparent presentation of results.
The suggested format for reporting results provides the company with an opportunity to
provide the following information.
· operations included within the operational boundaries of the inventory
· direct emissions
· indirect emissions and emissions attributable to exports of power and steam
· exports and imports of fossil fuel-derived CO2
· emission factors used to develop the inventory

Annex I suggests a format for reporting supporting information on biomass.

4.3 Data Quality

The calculation tools described in this report can be used to develop estimates for a number
of purposes. To a large extent, the purpose of the inventory will dictate the quality of the
data that is needed and the approach used to develop the inventory. In developing corporate
benchmarking data, for instance, it might be acceptable to use a generic emission factor for
coal burning, but a carbon trading program might require that emission estimates be based on
the carbon content of the specific coal being burned. The data quality requirements imposed
by the intended use of the inventory should be defined before the company begins the
inventory.

For most purposes, it will be acceptable for companies to estimate GHG emissions using
emission factors and corresponding “activity data” (e.g., amount of fuel used or miles
traveled). For most mills, the largest sources of GHG emissions are fossil fuel-fired
stationary combustion units. Fortunately, in most cases these emissions can be accurately
estimated because facilities usually have excellent records of the types and amounts of fossil
fuel being consumed, and CO2 emissions from these sources are directly related to fuel
carbon content as reflected in widely accepted emission factors.

For most other sources, however, the quality of GHG emission estimates is much lower,
sometimes because of inadequate activity data, but more often due to emission factors that
are based on very few data. Because of the importance of emission factors to the results of a

6
GHG inventory, these calculation tools include a table (Table 13) that companies are
encouraged to use to show the emission factors used to develop the inventory.

It can be expected that many more emissions measurements will be made in coming years
and improved emission factors will be developed reflecting these new data. Users of
emissions inventory data need to understand this process and the resulting impact it will have
on GHG inventory results. One can be certain that the quality of the estimates will improve
over time, but it is impossible to forecast whether, in the process, the estimates will be
adjusted upward or downward. The changes are expected to be relatively unimportant to the
GHG profile of most mills, however, because the largest sources of GHGs from most pulp
and paper mills, stationary fossil fuel combustion units, are well understood.

4.4 Units

Different countries use different units of measure. This can create considerable confusion
when emission factors and estimation techniques are applied internationally. To reduce the
confusion, the SI (metric) system is used throughout this report, but the Annexes contain
emission factors and other parameters in the units preferred by the authority or country
responsible for the information. Some of the important issues related to units of
measurement are highlighted here.

4.4.1 Units of Measure for Greenhouse Gases

Greenhouse gases are often compared on the basis of their estimated potential to cause global
warming. Factors have been developed, called Global Warming Potentials (GWPs), which
can be used to convert a quantity of non-CO2 greenhouse gas into an amount of CO2 with an
equivalent warming potential. Although the derivation of these factors involves a large
number of assumptions, GWPs are almost universally used to compare one greenhouse gas to
another. The GWP for CH4 is 21 so, from the standpoint of potential global warming, every
gram of CH4 is equivalent to 21 grams of CO2. The GWP for N2O is 310. The derivation of
these factors is explained elsewhere (IPCC 1996).2 When an emission estimate is the sum of
several GHGs expressed as the equivalent amount of CO2, the estimate is said to be in CO2-
equivalents, sometimes abbreviated as CO2e, CO2eq, or CO2-equiv. To convert CO2, CH4,
and N2O emissions into CO2-equivalents, multiply CH4 emissions by 21 and N2O emissions
by 310, then add both to CO2 emissions.

CO2-equivalents are also sometimes reported as the weight of the carbon in the CO2-
equivalents, usually reported in metric tons of carbon equivalents (MTCEs). MTCEs are
calculated by multiplying the weight of CO2-equivalents by 12/44, the weight fraction of
carbon in carbon dioxide.

For purposes of transparency and to avoid confusion, in these calculation tools greenhouse
gas quantities are usually reported in terms of the mass of the individual GHG, rather than

2
Recent research summarized in IPCC 2001 suggests that the GWP for CH4 should be higher (23) and that for
N2O should be lower (296) than the values previously recommended by IPCC (1996). However, the revised
GWPs have not been widely adopted at this time. Therefore, this report uses the GWP values recommended
by IPCC (1996) (21 for CH4 and 310 for N2O).

7
CO2-equivalents or MTCEs. In some cases, however, companies may find that it is
appropriate to use emission factors that are based on the combined emissions of several
GHGs expressed in terms of CO2-equivalents or MTCEs. This is acceptable provided it is
made clear in the results.

4.4.2 Heat Content of Fuels – GCV vs. NCV

Some countries measure fuel according to its gross calorific value (GCV) or higher heating
value (HHV), while other countries use net calorific value (NCV) or lower heating value
(LHV). The difference between these measures is that GCV and HHV include the heat of
condensation of water formed during combustion of the fuel. A common approximation is
that NCV is 95% of GCV for coal and oil and 90% of GCV for natural gas.

In this report, NCV (LHV) is used. In some cases, the emission factors have been converted
from GCV (HHV) units using the approximations described above. In the Annexes, energy-
related parameters are expressed in the units used by the authorities or countries that
developed the information. Except where noted, the Annexes use NCV (LHV).

5.0 IDENTIFYING CORE PULP AND PAPER OPERATIONS TO BE


INCLUDED WITHIN THE OPERATIONAL BOUNDARIES OF THE
INVENTORY

A number of operations are integral to pulp and papermaking. In this report, these operations
are defined as “core operations.” In many cases, companies will want to include these core
operation emissions within the operational boundaries of the inventory:
· direct emissions (i.e., company-owned emissions) from core operations that are on-site
(e.g., pulp mills, paper mills, power boilers)
· direct emissions from off-site core operations (e.g., company-owned harvesting
equipment and truck fleets)
· indirect emissions from on-site core operations that are wholly owned or partly owned by
another entity (including outsourced power islands)
· off-site indirect emissions related to power and steam imports
· on-site emissions attributable to power and steam exports

While companies may deviate from these boundary conditions based upon their particular
needs, it is important that the results be presented with a clear explanation of the operations
and emissions included within the boundaries of the inventory. A table is included in the
calculation tools (Table 10) which can be used to document the core operations within the
boundaries of the inventory and the ownership of those operations (i.e., the operational and
organizational boundaries of the inventory). Companies may use other formats to summarize
this information, but a description of the inventory boundaries and the core operations
included in the inventory should be included in the inventory results.

Considerations involved in setting operational boundary conditions for inventories and for
assigning ownership to emissions (i.e., setting organizational boundaries) are addressed in the
WRI/WBCSD GHG Protocol (WRI 2001a).

8
Examples of core operations that might be included in the operational boundaries of a GHG
inventory for an individual pulp and paper mill, as well as examples of several non-core
operations that would normally not be included (although they might be included in a
corporate inventory), are listed here.

Examples of core pulp and paper mill operations with the potential to emit greenhouse gases:
· power boilers, gas turbines, and other combustion devices producing steam and/or power
for the mill
· recovery furnaces and other devices burning spent pulping liquors
· incinerators
· lime kilns and calciners
· gas- or other fossil fuel-fired dryers (e.g., infrared dryers)
· anaerobic wastewater treatment or sludge digestion operations (usually included in the
boundaries of the inventory only if on-site or owned by the company)
· landfills used to dispose of mill wastes (usually included in the boundaries of the
inventory only if on-site or owned by the company)
· on-site vehicles and machinery
· harvesting equipment used to supply the mill (usually included in the boundaries of the
inventory only if owned by the company)
· trucks used to transport raw materials, products, or wastes for the mill (usually included
in the boundaries of the inventory only if owned by the company)

Examples of core pulp and paper mill operations that may be associated with the indirect
emission of greenhouse gases because they sometimes consume purchased power or steam:
· preparing virgin fiber (debarking, chipping and other woodyard operations, usually
included only if owned by the company)
· preparing recovered fiber - including deinking
· mechanical pulping
· chemical pulping
· semi-chemical pulping
· other chemical pulping processes
· chemical recovery operations
· pulp screening, thickening, washing
· virgin fiber bleaching and recovered fiber bleaching and brightening
· on-site production of chlorine dioxide and other bleaching chemicals
· paper and paperboard production, including stock cleaning and refining
· coating (including extrusion coating)
· trimming, roll wrapping, sheet cutting, etc.
· normal office and building operations for mill employees
· equipment for incoming process water treatment, waste treatment and disposal or
emissions control

Examples of operations that would not be core operations at pulp and paper mills:
· panel plants or wood chemical plants located at the mill site

9
· merchant power plants located adjacent to the mill whose primary business is selling
electricity
· converting operations that, at most mills, are not conducted on-site

Companies are encouraged to include within the operational boundaries of the inventory
outsourced core operations that significantly impact the facility’s GHG profile. The indirect
emissions from outsourced power islands, in particular, can be responsible for a large
fraction of the GHG emissions associated with a mill site.

The company must use its judgment to determine what is meant by an “on-site” operation. If
the company feels that its interpretation of the term “on-site” is important to understanding
the results of the inventory, this can be explained in the results. In most cases, however, the
ownership of the source (i.e., the organizational boundaries) will be much more important
than its location in determining how to handle its emissions in the inventory.

6.0 DETERMINING THE ORGANIZATIONAL BOUNDARIES OF THE


INVENTORY AND OWNERSHIP OF EMISSIONS

Ownership and control determine the organizational boundaries for GHG inventories. Direct
emissions, i.e., emissions from sources that the company owns or controls, are within the
company’s organizational boundaries and indirect emissions are outside those boundaries.
For purchases of electricity or steam, the situation is usually straightforward, because the
emissions are usually from sources outside the company’s ownership and control and are
therefore indirect. The reporting format suggested in these calculation tools divides
emissions along organizational boundary lines; i.e., into direct and indirect emissions.

Companies using these calculation tools, however, may encounter situations where inventory
boundaries include operations that are only partly owned by the company. Some or all of the
emissions from these sources might be “owned” by other entities and would therefore be
accounted for as indirect emissions. These situations often arise due to outsourcing and
partial ownership arrangements, which are becoming increasingly common in the pulp and
paper industry. Some examples include:
· combined heat and power operations where the mill is a partial owner of the generating
facilities
· power islands that are owned, at least in part, by other corporate entities
· wastewater treatment or landfill operations owned or operated by third parties
· mills where multiple parties share ownership of the facility

There are a large number of possible ownership arrangements, making it difficult to provide
specific instructions on how to derive organizational boundaries and determine ownership of
GHG emissions from these sources. Perhaps the most thorough discussion of methods for
determining the ownership of emissions from partly owned or partly controlled sources is
contained in the WRI/WBCSD GHG Protocol (WRI 2001a). Annex H to this report
contains an overview of the WRI/WBCSD guidelines. Companies performing inventories
that require allocating ownership of emissions from partly owned sources will want to obtain
the WRI/WBCSD GHG Protocol (WRI 2001a). The approach outlined in the GHG Protocol
can be summarized as follows.

10
Where the allocation of GHGs is specified contractually, this allocation is to be used.
Otherwise the WRI/WBCSD GHG Protocol suggests that allocation be done two ways,
according to control and according to equity share. The GHG Protocol recommends that the
results of both approaches be presented. The GHG Protocol emphasizes that organizational
boundaries developed for allocating GHGs “should be consistent with the organizational
boundaries which have been drawn up for financial reporting purposes” (WRI 2001a).

GHG allocation will often be governed by contractual arrangements and financial accounting
standards that will vary from situation to situation and between legal jurisdictions. For these
reasons, these calculation tools do not recommend a method for determining ownership of
emissions from partially owned sources. Companies should select methods that are most
appropriate for particular situations and should describe the methods in the inventory results.

7.0 MATERIALITY AND INSIGNIFICANT EMISSIONS

Greenhouse gas protocols generally allow facilities to ignore emissions that are so small that
they do not significantly impact the estimate of overall emissions. This concept of
“materiality” is drawn from financial reporting, where a material difference is sometimes
taken to be a discrepancy of more than 5% between reported and audited values (though this
is not an absolute standard) (Loretti et al. 2001). There is no generally accepted standard,
however, for materiality in GHG inventories (WRI 2001a; Loretti et al. 2001). Indeed, the
WRI/WBCSD GHG Protocol points out that “what is considered material will … depend on
the needs of users and the size of a company and its emissions sources.” It goes on to
observe that “the materiality of a source can only be established after it has been assessed.
This implies that [it] can be included in the inventory – even if it is just an estimate” (WRI
2001a).

These tools contain no specific recommendations on how to determine whether emissions are
so small that they can be ignored in preparing a GHG inventory, but they contain emission
factors and example calculations that may aid companies in deciding which emissions are
material and which are not. The decision on whether, or how, the estimates should be
reported must be left to the company. That decision may depend, in part, on a company’s
assessment of the quality of the data used to develop the estimate and the intended use of the
inventory results. In the results of the inventory, however, companies should identify the
materiality threshold that was used.

Table 1 has been developed from representative emission factors discussed in the Annexes to
this report. The information may assist companies in determining which sources must be
included in the inventory and which are so small that they can be ignored. Subsequent
sections of this report provide suggested default values for the factors. The factors in Table 1
clearly illustrate the importance of CO2 emissions from fossil fuel combustion. In most
cases, CH4 and N2O contribute relatively small quantities to a mill’s GHG emissions, even on
a CO2-equivalents basis. In addition, sources other than fossil fuel combustion will be
comparatively small. Other sections of this report contain example calculations that will be
helpful in identifying insignificant sources. The ultimate decision on which emissions to
include, however, must be made by the company.

11
Table 1. Emission Factor Ranges Useful in Identifying Significant and Insignificant Sources of GHGs
CH4 N2O Tables in Report Containing
Units Fossil-CO2 (CO2-equiv.)* (CO2-equiv.)* Default Values
Natural gas used in boilers kg CO2-equiv./TJ 56,100-57,000 13 – 357 31 – 620 2, 3
Residual oil used in boilers kg CO2-equiv./TJ 76,200 – 78,000 13 – 63 93 – 1550 2, 3
Q
Coal used in boilers kg CO2-equiv./TJ 92,900-126,000 15 – 294 155 – 29,800 2, 3
Bark and wood waste fuel kg CO2-equiv./TJ 0 <21 – 860 <310 – 8060 6
Black liquor kg CO2-equiv./TJ 0 42 – 630 1550 6
Lime kilns kg CO2-equiv./TJ Depends on fuel 21 – 57 0y 2, 4
l
Lime calciners kg CO2-equiv./TJ Depends on fuel 21 – 57 1550 2, 4
Pulp mill make-up CaCO3 kg CO2/ t CaCO3 440 0 0 5
Pulp mill make-up Na2CO3 kg CO2/ t Na2CO3 415 0 0 5
Diesel fuel used in vehicles kg CO2 equiv./TJ 74,000 – 75,300 82 – 231 620 – 9770 7, 8
Gasoline in non-road mobile sources kg CO2 equiv./TJ 69,300 – 75,300 84 – 30,900 93 – 2580 8
and machinery – 4-stroke engines
Gasoline in non-road mobile sources kg CO2 equiv./TJ 69,300 – 75,300 9,860 – 162,000 124 – 861 8
and machinery – 2-stroke engines
Anaerobic wastewater treatment kg CO2 eq/ kg COD treated 0 5.25 h 0 Equations 6 and 7
Mill solid waste landfills kg CO2 eq/ dry ton solid 0 3,500 ¥ 0 Equations 3,4,5 and Table 9
waste
* CO2-equivalents are calculated from IPCC Global Warming Potentials (CH4 = 21 and N2O = 310).
Q
Reported N2O emission factors greater than 1500 kg CO2 eq./TJ are generally limited to fluidized bed boilers
y
IPCC information suggests N2O is not likely to be formed in lime kilns in significant amounts.
l
Amounts of N2O, if any, formed in calciners is not known, so the largest factor for fuels normally used in kilns is shown here.
h
Assumes no capture of gas from the treatment plant.
¥
Assumes that 50% of landfilled waste is degradable organic carbon, 50% of the degradable organic carbon degrades to gas, 50% of the carbon in the gas is
contained in methane, none of the methane is oxidized in the landfill cover or captured, and all is released in the same year that the waste is landfilled.
This method is used here only to generate an emission factor for considering whether to include this source in the inventory. More refined methods,
which will normally yield lower estimates of emissions, are explained in the calculation tools.
In the results of the inventory, companies should identify those situations where emissions
have been estimated to be too small to materially impact the inventory results. The reporting
format allows these situations to be identified by reporting these releases as “non-material”
or “NM” in the results. Companies should also indicate in the results the criteria used to
decide whether emissions are non-material. A footnote can be added, for instance, indicating
that the emissions are non-material because they represent less than a certain percent of the
mill’s or company’s direct emissions.

8.0 GREENHOUSE GAS EMISSIONS FROM STATIONARY FOSSIL FUEL


COMBUSTION

8.1 Carbon Dioxide

Carbon dioxide emissions from stationary fossil fuel combustion represent the majority of
GHG emissions for most pulp and paper mills. Emissions of CO2 are estimated from the
carbon content of, or emission factors for, all fossil fuels being burned. In some cases, a
correction (i.e., a reduction) is made for unburned carbon. Companies can use data from one
of the following sources, with the preferred sources listed first.
· data on the specific fuels and boilers being used at the mill
· the most appropriate data recommended by the national authorities
· the most appropriate data available from other sources

CO2 emission factors and information on fossil fuel carbon content and unburned carbon are
available from most national authorities and a variety of existing protocols. Some of this
information is included in Annex A to this report. In cases where the Annex contains both
country-specific information and IPCC information, the country-specific information may be
preferred because it may reflect country-specific data on fuel carbon content. The IPCC
generic emission factors are shown in Table 2.

To correct CO2 emissions estimates for unburned carbon, IPPC recommends default
correction factors of 0.98 for coal, 0.99 for oil and oil products, 0.995 for gas, and 0.99 for
peat (non-household use combustion) (IPCC 1997c). The IPCC emission factors in Table 2
are presented as corrected for unburned carbon, and as uncorrected. IPCC points out,
however, that in the case of coal, unburned carbon can be much higher than the default
values and cites an Australian study of coal-fired boilers wherein unburned carbon ranged
from 1 to 12% of the carbon fed to the boiler.

In many cases, it will be possible to estimate the total CO2 emissions from all sources
burning a single fossil fuel without estimating the emissions from each unit separately. For
instance, if a mill is burning natural gas in a boiler for steam production and in several
infrared dryers, the CO2 emissions from natural gas burning can be estimated from the total
gas used for both purposes.

If the mill exports fossil fuel-derived CO2, for instance to an adjacent precipitated calcium
carbonate (PCC) plant, these exports should be not be included in the emissions estimates
because this CO2 is not emitted by the mill. A separate line is included in the results table
(Table 12) to report exports of fossil fuel-derived CO2.

13
Table 2. IPCC Default CO2 Emission Factors for Fossil Fuels (after IPCC 1997b)
Fossil Fuel Uncorrected Emission Factor Corrected Emission Factor
kg CO2/TJ* kg CO2/TJ
Crude oil 73,300 72,600
Gasoline 69,300 68,600
Kerosene 71,900 71,200
Diesel oil 74,100 73,400
Residual fuel oil 77,400 76,600
LPG 63,100 62,500
Petroleum coke 100,800 99,800
Anthracite coal 98,300 96,300
Bituminous coal 94,600 92,700
Sub-bituminous coal 96,100 94,200
Lignite 101,200 99,200
Peat 106,000 104,900
Natural gas 56,100 55,900
* These factors assume no unburned carbon. To account for unburned carbon, IPCC
suggests multiplying by these default factors: coal = 0.98, oil = 0.99, and
gas = 0.995.

Municipal solid waste (MSW) or materials derived from MSW are sometimes used as a fuel.
MSW contains a combination of fossil carbon (primarily in plastics) and climate-neutral
biomass carbon (primarily in paper and food waste). IPCC recommends that the composition
of the MSW (i.e., its fossil carbon content) be used to estimate emissions of fossil CO2.
Where no other data are available, IPCC recommends assuming that 16% of the wet weight
of MSW is fossil carbon and 5% of the fossil carbon is unburned (IPCC 2000a). These
combined assumptions yield a fossil-CO2 emission factor of 557 kg CO2/wet tonne MSW
burned.

8.2 Methane and Nitrous Oxide

Methane and nitrous oxide emissions from fossil fuel combustion are usually very small
compared to CO2 emissions. Indeed, some inventory protocols, including the WRI/WBCSD
GHG Protocol, do not address CH4 and N2O from fossil fuel combustion. Because some
inventories include CH4 and N2O, however, they are addressed in these calculation tools.

Companies will often be able to use the data in Table 1 to demonstrate that emissions of CH4
and N2O from fossil fuel combustion are insignificant compared to CO2 emissions. In other
cases, a single emission factor might be available that includes fossil-CO2, CH4, and N2O

14
emissions expressed as CO2 equivalents. In this case, the company does not need to report
the three gases separately.

Estimating CH4 and N2O emissions will usually involve selecting the emission factors best
suited to the fuels being burned and the type of combustion unit. For normal fossil fuel-fired
boilers, recommended emission factors follow the following order of preference.
· data on the specific fuels and boilers being used at the mill
· the most appropriate data recommended by the national authorities
· the most appropriate data available from other sources

A number of existing protocols and most national authorities publish factors for estimating
emissions of CH4 and N2O from fossil fuel-fired boilers. Many of these emission factors are
listed in Annex A to these tools.

IPCC provides Tier 1 and Tier 2 emission factors for methane and nitrous oxide. IPCC’s
Tier 2 emission factors, shown in Table 3, are usually preferred because they are more
specific to fuel type and combustion device. Tier 2 emission factors, however, are based on
uncontrolled emissions. This is unimportant for methane because emission control devices
have little impact on methane emissions (IPCC 1997c). Nitrous oxide emissions can be
impacted by control devices, but the data are very limited (IPCC 1997c). Where N2O
emissions are important to inventory results, companies may want to develop emissions data.
In most cases, however, the difference between controlled and uncontrolled emissions is
expected to have little effect on total GHG emissions. Therefore, companies will probably
want to use the Tier 2 emission factors shown in Table 3 unless other factors are available
that are more suited to individual mill circumstances. The table presents IPCC’s Tier 2
emission factors for the fossil fuels and combustion devices of most interest to the pulp and
paper industry.

Methane and nitrous oxide emission factors for stationary internal combustion engines (e.g.,
those used to drive emergency generators or turbines) fired with diesel or gasoline can be
approximated by the factors for non-road mobile sources shown in Table 8.

Fossil fuels can be used in a number of different combustion units common in the forest
products industry. Some specific recommendations for estimating CH4 and N2O emissions
from certain types of units found exclusively or primarily in the forest products industry are
shown below.

Recovery furnaces – In most cases, only small amounts of fossil fuel are burned in recovery
furnaces. The CO2 releases from the fossil fuel can be estimated using the methods described
in Section 8.1. Where large amounts of fossil fuels are being burned (i.e., they represent a
major source of fuel on an ongoing basis), best professional judgment will be required to
select the most appropriate emission factors for CH4 and N2O. In the vast majority of cases,
however, the small amounts of fossil fuel used in the recovery furnace can be included in the
firing rate used to estimate CH4 and N2O from liquor burning. The CH4 and N2O emissions
factors for recovery furnaces are included in the Section 11.2.

15
Table 3. IPCC Tier 2 Uncontrolled CH4 and N2O Emission Factors
for Industrial Boilers (IPCC 1997c)
Fuel Technology Configuration kg CH4/TJ kg N2O/TJ
Bituminous coal Overfeed stoker 1.0 1.6
boilers
Sub-bituminous coal Overfeed stoker 1.0 1.6
boilers
Bituminous coal Underfeed stoker 14 1.6
boilers
Sub-bituminous coal Underfeed stoker 14 1.6
boilers
Bituminous coal Pulverized Dry bottom, wall fired 0.7 1.6
Bituminous coal Pulverized Dry bottom, tang. fired 0.7 0.5
Bituminous coal Pulverized Wet bottom 0.9 1.6
Bituminous coal Spreader Stoker 1.0 1.6
Bituminous coal Fluidized bed Circulating or bubbling 1.0 96
Sub-bituminous coal Fluidized bed Circulating or bubbling 1.0 96
Anthracite 10* 1.4*
Residual oil 3.0 0.3
Distillate oil 0.2 0.4
Natural gas Boilers 1.4 0.1*
Natural gas Turbines 0.6 0.1*
Natural gas Int. comb. engine 2-cycle lean burn 17 0.1*
Natural gas Int. comb. engine 4-cycle lean burn 13 0.1*
Natural gas Int. comb. engine 4-cycle rich burn 2.9 0.1*
* These are IPCC Tier 1 generic emission factors for coal and natural gas. Tier 2 emission factors are
not available.

Combination fuel-fired boilers burning biomass and fossil fuels – The CO2 releases from
fossil fuel used in these boilers can be estimated using the methods described in Section 8.1.
In most cases, the combustion conditions in these boilers are more like those in biomass-fired
boilers than fossil fuel-fired boilers. Therefore, unless data are available from site-specific
testing on other similar boilers burning a comparable mix of fuels, it is recommended that the
emissions from combination fuel-fired boilers be estimated from the total heat input to the
boiler and CH4 and N2O emission factors for biomass. These emission factors are
summarized in Section 11.2.

Kraft mill lime kilns and calciners – The emissions from lime kilns and calciners are unique
enough to warrant separate discussion.

16
Gas-fired infrared dryers, incinerators, and other miscellaneous pulp and paper sources –
The CO2 releases from the fossil fuel used in these units can be estimated using the methods
described in Section 8.1. Lacking site-specific information, companies will either have to
assume that the methane and nitrous oxide emissions from these sources are negligible
(based on the amounts of fuel burned) or use the emission factors developed for similar fuels
burned in other operations. The data in Table 1 suggest that it should be relatively simple for
many companies to document that these sources of CH4 and N2O are so small that they can
be ignored in the inventory. The fossil CO2 from these sources can be estimated directly
from the carbon content of the fuel or CO2 emission factors, using the same methods as for
other stationary combustion units.

Example Calculation: CO2, CH4, and N2O emissions from natural gas use at a small mill.

A mill uses natural gas in a small boiler and in several infrared dryers. The mill’s records
indicate that over a year’s time, it used 17 million standard cubic meters of natural gas. The
mill decides to estimate the emissions from overall natural gas consumption instead of
attempting to separate boiler emissions from the infrared dryer emissions. The mill does not
know the carbon content of its gas supply, but the IPCC emission factor is 55.9 metric tons
CO2/TJ (after correcting for 0.5% unoxidized carbon). The mill uses the CH4 and N2O
emission factors from Table 3 (1.4 kg CH4/TJ and 0.1 kg N2O/TJ). The mill estimates the
heating value of the natural gas to be 52 TJ/kiloton and the density to be 0.8 kg/standard
cubic meter. The annual emissions are estimated as follows.

CO2 emissions:
· (17 x 106 m3 gas/y) x (0.8 kg/m3) = 13.6 x 106 kg gas/y = 13.6 kiloton gas/y
· (13.6 kiloton gas/y) x (52 TJ/kiloton) = 707 TJ/y
· (707 TJ/y) x (55.9 t CO2/TJ) = 39,500 t CO2/y

CH4 emissions
· (707 TJ NCV/y) x (1.4 kg CH4/TJ NCV) = 990 kg CH4/y = 0.99 t CH4/y
Using the IPCC GWP of 21, this is equal to 21 t CO2-eq./y

N2O emissions
· (707 TJ NCV/y) x (0.1 kg N2O/TJ NCV) = 70.7 kg N2O/y = 0.07 t N2O/y
Using the IPCC GWP of 310, this is equal to 22 t CO2-eq./y

Total GHG emissions = 39,500 + 21 + 22 = 39,500 t CO2-equivalents/y

On a CO2-equivalents basis, CH4 and N2O emissions are barely 0.1% of CO2 emissions.
These emissions would be very small even if the emission factors were several times larger.
Because of this, the company might decide against including CH4 and N2O estimates in the
results, and instead indicate in the results that the estimates demonstrate that the emissions
are not material to the results of the inventory because they are less than 0.1% of CO2
emissions.

17
Example Calculation: CO2, CH4, and N2O emissions from a large dry-bottom, wall fired
boiler burning pulverized bituminous coal.

The boiler produces 350,000 kg steam per hour (about 770,000 pounds/hr). Over a year’s
time, the mill’s records indicate that the boiler consumed 336,000 Mg (370,000 short tons) of
coal having a higher heating value, on average, of 13,000 Btu HHV/lb.

Case 1: CO2 emissions based on carbon content of fuel

The mill has information on the carbon content of the coal being burned in the boiler (80.1%
carbon, by weight). The mill decides that the default IPCC correction for unburned carbon in
coal-fired boilers (2% unburned carbon) is appropriate. The mill decides to use the Tier 2
IPCC emission factors for CH4 and N2O from Table 3. The IPCC Tier 2 emission factors for
dry bottom, wall fired boilers burning pulverized bituminous coal are 0.7 kg CH4/TJ NCV
and 1.6 kg N2O/TJ NCV. The mill applies the usual assumption that the NCV (or LHV) for
coal is 5% lower than the GCV (or HHV). The annual emissions of CO2, CH4, and N2O are
estimated as follows.

CO2 emissions:
· (336,000 Mg/y coal) x (0.801 Mg carbon / Mg coal) x ( 0.98 Mg carbon burned) x
(44 Mg CO2 / 12 Mg carbon) = 967,000 Mg CO2/yr or 967 x 103 t CO2/yr

CH4 emissions:
· 370,000 short tons coal/y = 740 x 106 pounds/y
· (740 x 106 pounds/y) x (13,000 Btu HHV/pound) = 9.62 x 1012 Btu HHV/y
· for coal, LHV is 0.95 times HHV (see Section 4.4.2)
· (9.62 x 1012 Btu HHV/y) x (0.95 to correct to LHV) = 9.14 x 1012 Btu LHV/y
· (9.14 x 1012 Btu LHV/y) x (1055 J/Btu) = 9.64 x 1015 J NCV/y = 9.64 x 103 TJ NCV/y
· CH4 emissions = (9.64 x 103 TJ NCV/y) x (0.7 kg CH4/TJ NCV) = 6.75 x 103 kg CH4/y
or 6.75 t CH4/y
Using the IPCC GWP of 21 for CH4, this equates to 142 t CO2-eq./y

N2O emissions:
· N2O emissions = (9.64 x 103 TJ NCV/y) x (1.6 kg N2O/TJ NCV) = 15.4 t N2O/yr
Using the IPCC GWP of 310 for N2O, is to 4780 t CO2-eq./y

Total GHG emissions = 967,000 + 142 + 4,780 = 972,000 t CO2-equivalents/y

Compared to its CO2 emissions, emissions of CH4 and N2O from this boiler are very small.
CO2 emissions = 967,000 t CO2/y
CH4 emissions = 142 t CO2-eq./y or 0.015% of CO2 emissions
N2O emissions = 4,780 t CO2-eq./y or only about 0.5% of CO2 emissions

18
Case 2: CO2 emissions based on emission factors

In this case the mill does not have information on the carbon content of the coal being burned
in the boiler. The IPCC default Tier 1 emission factor for CO2 is 94.6 t CO2/TJ NCV. The
mill decides that the default IPCC correction for unburned carbon in coal-fired boilers (2%
unburned carbon) is appropriate.

CO2 emissions:
· 370,000 short tons coal/y = 740 x 106 pounds/y
· (740 x 106 pounds/y) x (13,000 Btu HHV/pound) = 9.62 x 1012 Btu HHV/y
· for coal, LHV is 0.95 times HHV (see Section 4.4.2)
· (9.62 x 1012 Btu HHV/y) x (0.95 to correct to LHV) = 9.14 x 1012 Btu LHV/y
· (9.14 x 1012 Btu LHV/y) x (1055 J/Btu) = 9.64 x 1015 J NCV/y = 9.64 x 103 TJ NCV/y
· uncorrected CO2 emissions = (9.64 x 103 TJ NCV/y) x (94.6 t CO2 /TJ NCV) = 912 x 103
t CO2/y
· CO2 emissions corrected for 2% unburned carbon = 894 x 103 t CO2/y

Methane and nitrous oxide emissions are calculated as in Case 1.

CO2 emissions = 894,000 t CO2/y


CH4 emissions = 142 t CO2-eq./y or 0.016% of CO2 emissions
N2O emissions = 4,780 t CO2-eq./y or only about 0.5% of CO2 emissions

As in the previous example, this comparison suggests that it may be relatively simple for
many mills to document that emissions of CH4 and N2O from fossil fuel-fired boilers are not
material to the results of their inventory.

9.0 EMISSIONS FROM KRAFT MILL LIME KILNS AND CALCINERS

Fossil-CO2 emissions from kraft mill lime kilns and calciners are estimated using the same
approach as used for other stationary fossil fuel combustion devices – by determining how
much fossil fuel is used in the kiln and estimating emissions from information on the fuel
carbon content or emission factors. These CO2 emissions are reported together with other
fossil fuel-related CO2 emissions.

Although CO2 is also liberated from the CaCO3 burned in the kiln or calciner, the carbon
released from CaCO3 is climate-neutral biomass carbon that originates in wood and should
not be included in GHG emissions totals. Companies wanting to prepare inventory reports
that are consistent with the requirements of the WRI/WBCSD GHG Protocol should report
these climate-neutral biomass-related emissions as “supporting information” using Annex I.
The movement of carbon in kraft pulp mills and the reasons for differentiating biomass- from
fossil-CO2 emissions from lime kilns are explored in detail in Annex B to this report and in a
paper by Miner and Upton (2002).

There are very few data on CH4 and N2O emissions from kraft mill lime kilns and calciners.
This review uncovered data from only three lime kilns sampled in the early 1980s (NCASI
1981). These data suggest an emission factor of 2.7 kg CH4/TJ. For commercial lime kilns,

19
IPCC suggests emission factors of 1.0 and 1.1 kg CH4/TJ for oil-fired and gas-fired lime
kilns, respectively. The IPCC factors are for commercial lime kilns, however, and may not
be appropriate for kraft mill lime kilns. Table 1 illustrates that for fossil fuel-fired kilns or
calciners, CH4 emissions will be very small compared to the CO2 from fossil fuel.

No data were found for N2O from lime kilns or calciners, but the temperatures in rotary lime
kilns appear to be too high to allow significant generation of N2O (see Annex B for more
information). It is reasonable, therefore, to assume that N2O emissions from rotary lime kilns
are negligible. The temperatures in calciners appear to be more amenable to N2O generation
(see Annex B). Given the range of N2O emission factors for oil and gas, however, it seems
likely that N2O emissions will be small relative to fossil-CO2 emissions from fossil fuel-fired
calciners.

The emission factors suggested for kraft mill lime kilns and calciners are summarized in
Table 4.

Table 4. Emission Factors for Kraft Mill Lime Kilns and Calciners
Emissions, kg/TJ
Fuel Kraft mill lime kilns Kraft mill calciners
CO2 CH4 N2O CO2 CH4 N2O
Residual oil 76,600* 2.7 q 0c 76,600* 2.7 q 0.3 f
Distillate oil 73,400* 2.7 q 0c 73,400* 2.7 q 0.4 f
Natural gas 55,900* 2.7 q 0c 55,900* 2.7 q 0.1 f
Biogas 0 2.7 q 0c 0 2.7 q 0.1 f
*
from Table 2, corrected for unburned carbon
q
from NCASI 1981
c
based on IPCC description of temperatures giving rise to N2O emissions
f
from Table 3

At a number of mills around the world, stack gas from lime kilns or calciners is piped to
adjacent precipitated calcium carbonate (PCC) plants for use as a raw material (PCC is
sometimes used as an inorganic filler or coating material in paper and paperboard products).
Because these so-called “satellite” PCC plants are not present at most mills, and because they
are almost always owned by a separate company, they are not considered core operations in
these calculation tools.

A separate line is included in the inventory results to show the amounts of fossil fuel CO2
being exported to PCC plants. These exports of fossil fuel-derived CO2 should not be
included in the estimates of GHG emissions because they are not emitted by the mill. If the
mill also wants to show the amounts of climate-neutral biomass-derived CO2 that are
exported with lime kiln stack gas, it can include this as supporting information (see Annex I).

A mill may sometimes export CO2 to the PCC plant when the PCC plant is not operating.
These calculation tools do not require that the exports be corrected to account for these

20
periods because the mill no longer owns the exported CO2 and cannot control whether the
PCC plant uses it. Presumably, if the PCC plant were to conduct an inventory, any unused
fossil-CO2 received from the mill would be shown as a direct emission in its inventory.
Exports of fossil fuel CO2 to PCC plants are shown in the results table (Table 12). This table
is also used to show any imports of CO2 to the mill, for use in neutralization, for instance.

These draft calculation tools do not address the ultimate fate of exported CO2 or the
ownership of any CO2 that is ultimately emitted from the PCC plant, as both issues are
outside the boundary of core pulp and paper operations. Nor does it address the question of
how much of this carbon is sequestered as a result of being converted to PCC. Ultimately,
however, most of the carbon in PCC is sequestered in landfilled paper, landfilled residuals
from deinking mills, or in ash from burning used paper.

Example Calculation: GHG emissions from a natural gas-fired lime kiln.

A 1000 ton/day kraft mill has a single gas-fired lime kiln. The mill’s records indicate that it
used 28.6 x 106 pounds of gas last year with a typical heat content of 21,000 Btu HHV/lb and
a density of 0.77 kg/m3. The IPCC CO2 emission factor for natural gas from boilers can be
used for lime kilns since the CO2 emissions are a function only of gas composition. The
IPCC CO2 emission factor for natural gas is 55.9 t CO2/TJ (after correcting for 0.5%
unoxidized carbon). For CH4, the mill decides to use the only available emission factor for
kraft mill lime kilns (2.7 kg CH4/TJ) and assumes that N2O emissions are negligible based on
the IPCC discussion of temperatures needed to generate N2O. The kiln’s GHG emissions
are estimated as follows.

CO2 emissions:
· 28.6 x 106 pounds gas/y x 21,000 Btu HHV/lb = 601 x 109 Btu HHV/y
· for natural gas, LHV is 0.9 times HHV (see Section 4.4.2)
· 601 x 109 Btu HHV/y x 0.9 (to convert to LHV) = 541 x 109 Btu LHV/y
· 541 x 109 Btu/y x (1.055 x 10-6 GJ/Btu) = 570,000 GJ/y = 570 TJ/y
· 570 TJ/y x 55.9 t CO2/TJ = 31,900 t CO2/y

CH4 emissions:
· 570 TJ/y x 2.7 kg CH4/TJ = 1540 kg CH4/yr
Using the IPCC CO2 equivalency factor of 21, this equals 32 t CO2-equivalents. This is a
very small number compared to the CO2 emissions (31,900 t). In addition, the estimate is
based on a small and old data set. Consequently, in the inventory results the mill might
decide to report that the CH4 emissions from this source are non-material to the inventory.

N2O emissions:
· As discussed above and in more detail in Annex B, IPCC’s analysis of the temperatures
needed to form N2O in combustion processes suggests that it is unlikely that significant
amounts of N2O would be formed in lime kilns. The mill would probably decide to note this
in the inventory results.

Total GHG emissions = 31,900 + 32 + 0 = 31,900 CO2-equivalents/y

21
10.0 CARBON DIOXIDE EMISSIONS FROM MAKE-UP CHEMICALS

Although losses of sodium and calcium from the recovery system are usually made up using
non-carbonate chemicals, small amounts of CaCO3 and Na2CO3 are sometimes used. The
carbon contained in these chemicals is usually of fossil origin, although in some cases (e.g.,
Na2CO3 purchased from soda-based semi-chem mills) it can be derived from biomass. In
these calculation tools, it is assumed that the carbon in these make-up chemicals escapes as
CO2 from the lime kiln or recovery furnace. These emissions are estimated by assuming that
all of the carbon in CaCO3 and Na2CO3 used in the recovery and causticizing areas is
released to the atmosphere. The amounts are usually small enough that, under normal
circumstances, it is reasonable to use either mill purchasing records or industry norms to
develop the estimates. If the carbon in make-up chemicals is biomass in origin (an
uncommon situation), the CO2 released from it is not considered a GHG emission and, in
these cases, this source does not need to be addressed (although it can be included as
Supporting Information on Biomass in Annex I).

The conversion factors for estimating fossil-CO2 releases from the use of carbonate-based
make-up chemicals in the pulp mill are shown in Table 5.

Table 5. Emissions from Calcium- and Sodium-Carbonate Make-up in the Pulp Mill*
Emissions
Pulp mill make-up CaCO3 440 kg CO2/ t CaCO3
Pulp mill make-up Na2CO3 415 kg CO2/ t Na2CO3
* If the carbonate is derived from biomass, GHG emissions are
zero.

It is important to note that calcium make-up is required because of losses from the
causticizing area, most of which are in the form of calcium carbonate. This lost material is
usually landfilled, thereby sequestering the carbon contained in the calcium carbonate.
Because the default method in these calculation tools does not consider this loss of carbon
from the system, the estimated CO2 emissions from make-up calcium carbonate will be
higher than actual emissions. Where these emissions are significant, companies may want to
perform the more detailed analyses required to correct the emissions estimates to account for
the carbon that leaves the causticizing area in calcium carbonate.

Example Calculation: Kraft mill using CaCO3 for make up at the lime kiln.

A 2000 tpd kraft mill determined from mill records that it uses about 7000 t (7700 short tons)
CaCO3 a year as make-up in the causticizing area (make-up rate of about 2% for this mill).
This CaCO3 is from a source where carbonate would be expected to be fossil (not biomass) in
origin. The emissions are estimated as follows.

· (7000 t CaCO3/y ) x 440 kg CO2/t CaCO3 = 3,080,000 kg CO2/y = 3080 t CO2/y

22
11.0 EMISSIONS FROM STATIONARY COMBUSTION OF BIOMASS FUELS

11.1 Climate-Neutral Carbon Dioxide Emissions from Burning Biomass Fuels

Many pulp and paper mills generate more than half of their energy needs from climate-
neutral biomass fuels recovered from the industry’s waste and process streams. Energy-rich
biomass – derived from wood chips, bark, sawdust, and pulping liquors recovered from the
harvesting and manufacturing processes – is formed when atmospheric carbon dioxide is
sequestered by trees during growth and transformed into organic carbon substances. When
these biomass fuels are burned, the CO2 emitted during the manufacturing and combustion
processes is the atmospheric carbon dioxide that was sequestered during growth of the tree;
hence, there is no net contribution to the atmospheric CO2 level. This carbon cycle is a
closed loop. New tree growth keeps absorbing atmospheric carbon dioxide and maintains the
cycle.

Any increases or decreases in the amount of carbon sequestered by the forests are accounted
for in the comprehensive forest accounting system. This is the approach generally prescribed
for national inventories by the United Nations Framework Convention on Climate Change.
Most international protocols, including that of the Intergovernmental Panel on Climate
Change (IPCC), have adopted the convention set out by the United Nations. The IPCC has
stated that emissions from biomass do not add to atmospheric concentrations of carbon
dioxide (IPCC 1997a, c).

Therefore, in keeping with well-established practices, the greenhouse gas inventory results
generated using these calculation tools do not include CO2 emissions from biomass burning.
The WRI/WBCSD GHG Protocol and some national reporting schemes, however, require
that these climate-neutral emissions be estimated and reported as “supporting information.”
Annex I contains information that will assist companies that want to comply with these
requirements.

IPCC provides the following list of biomass fuels (IPCC 1997a, c):
· wood and wood waste (although biogas from wood waste and other biomass is not
specifically listed by IPCC, it clearly falls within the general definition of biomass)
· charcoal
· dung
· agricultural residues and wastes
· municipal and industrial wastes (where the organic material is biological in origin) (this
would include wastewater treatment sludges from pulp and paper mills)
· bagasse
· bio-alcohol
· black liquor
· landfill gas
· sludge gas

CO2 emissions from peat burning are usually considered to be GHGs and they are included in
the emissions from fossil fuel burning (Table 2).

23
11.2 Methane and Nitrous Oxide Emissions from Burning Biomass Fuels

Although CO2 from biomass burning is almost universally excluded from GHG inventories,
CH4 and N2O from biomass burning are sometimes included. Calculation tools are provided,
therefore, to assist in estimating these emissions.

If a company has reliable site-specific data allowing it to estimate CH4 and N2O emissions
from biomass combustion, it should use these data. Otherwise, it will be necessary to use the
most appropriate emission factors available. Unfortunately, there are few data on CH4 and
N2O emissions from biomass boilers and recovery furnaces. Some of the available emission
factors are shown here and discussed in more detail in Annex C

IPCC’s Revised 1996 Inventory Guidelines used emission factors developed by EPA. This is
also true for a number of countries. EPA has since revised these factors. Even the updated
EPA factors, however, are based on very few data. Table 6 provides a summary of the
available information on methane and nitrous oxide emissions from biomass boilers. The
many individual country factors that are based on IPCC or EPA factors are not shown. The
variability in the data reflects the many different types and ages of boilers tested, operating
conditions, control equipment, and fuel characteristics.

Where an emission factor shown in the table was developed for circumstances that match
conditions at a mill, the company may want to select that emission factor for estimating mill
emissions. For instance, companies with fluidized bed boilers may want to use the Fortum
emission factors because they were developed on fluidized bed boilers while the other
emission factors were developed on stoker boilers or on boilers of an unspecified design. In
many cases, however, because of the ranges in emission factors and the limited ability, at this
time, to match emission factors to boiler designs, operating conditions, and fuels, it is
reasonable to use the median emission factors shown in the table. These median emission
factors fall within the ranges cited in the CORINAIR emissions inventory (ranges also shown
in the table) (EEA 1999).

Combination fuel-fired boilers burning biomass and fossil fuels – In most cases, the
combustion conditions in these boilers are more like those in biomass-fired boilers than fossil
fuel-fired boilers. Methane and nitrous oxide emissions are often more directly related to
combustion conditions than to fuel type. Because of the high moisture content of most
wood-based fuels, a reasonable default is to assume that the combustion conditions in
combination fuel-fired boilers reflect the impact of the wood wastes. Therefore, unless data
are available from site-specific testing on other similar boilers burning a comparable mix of
fuels, or unless the combustion conditions in the combination fuel-fired boiler are more like
fossil fuel-fired boilers than wood waste boilers, it is recommended that the emissions from
combination fuel-fired boilers be estimated from the total heat input to the boiler and CH4
and N2O emission factors for biomass.

24
Table 6. Emission Factors for CH4 and N2O from Biomass Combustion
kg kg
Emission Factor Description CH4/TJ N2O/TJ Reference
Wood waste-fired boilers
Wood, wood waste, and other biomass and 30 4 Tier 1 – IPCC 1997c
wastes
Uncontrolled emissions from wood-fired 15 - Tier 2 – IPCC 1997c
stoker boilers
Average for wood residue combustion 9.5* 5.9* USEPA 2001
Average for circulating fluidized bed boilers 1 8.8 Fortum 2001
burning peat or bark
Average for bubbling fluidized bed boilers 2** <2 Fortum 2001
burning peat or bark
Pre-1980 wood residue-fired stoker boilers 8.2* - NCASI 1980
sampled ahead of control devices
Pre-1980 wood residue-fired stoker boilers 2.7* - NCASI 1985
sampled after wet scrubbers
Wood fired boiler 41l 3.1l JPA 2002
Wood as fuel 24 l 3.4 l AEA Tech. 2001
Wood waste 30 5 Swedish EPA 2001
Median emission factors for wood waste 12 4
1 - 40 1.4 – 75 EEA 1999
Recovery furnaces
Recovery furnace <1 <1 Fortum 2001
Recovery furnace – black liquor 2.5 W - JPA 2002
Black Liquor 30 5 Swedish EPA 2001
Median emission factors for black liquor 2.5 2
1 –17.7 1 – 21.4 EEA 1999
* converted from GCV to NCV assuming a 5% difference
** excludes one very high number associated with low oxygen-high carbon monoxide conditions
l
based on heat content of 20 GJ/t dry solids
W
based on liquor heat content of 13.3 GJ/t dry solids

25
Example Calculation: Mill with a bark boiler.

A mill has a 250,000 kg steam/hour (550,000 pound/hr) circulating fluidized bed (CFB) bark
boiler. In a year, the boiler burns approximately 6.9 x 106 GJ of bark and 0.8 x 106 GJ of
residual fuel oil. Because the boiler receives supplemental fossil fuel, it is necessary to
estimate the CO2 from the fossil fuel use and the CH4 and N2O emissions based on the total
firing rate. The mill decides to use the IPCC emission factor for residual oil (76.6 t CO2/TJ,
after correcting for 1% unoxidized carbon) and to estimate CH4 and N2O emissions based on
the total firing rate and the emission factors developed by Fortum on CFB boilers. The
average emission factors found by Fortum, shown in Table 6, are 1 kg CH4/TJ and 8.8 kg
N2O/TJ.

CO2 emissions from fossil fuel:


· (0.8 x 106 GJ/y) = (0.8 x 103 TJ/y)
· (0.8 x 103 TJ/y) x (76.6 t CO2/TJ) = 61,300 t CO2/y

CH4 emissions:
· total heat input = (6.9 x 106 GJ/y) + (0.8 x 106 GJ/y) = 7.7 x 106 GJ/y = 7.7 x 103 TJ/y
· 7.7 x 103 TJ/y x 1 kg CH4/TJ = 7,700 kg CH4/y = 7.7 t CH4/y
Using the IPCC warming potential of 21, this equates to 162 t CO2-eq./y.

N2O emissions:
· total heat input = 7.7 x 103 TJ/y
· 7.7 x 103 TJ/y x 8.8 kg N2O/TJ = 67,800 kg N2O/y = 67.8 t N2O/y
Using the IPCC warming potential of 310, this equates to 21,000 t CO2-eq./y

Total CO2 equivalents emitted = 61,300 + 162 + 21,000 = 82,500 t CO2-equivalents/y

12.0 EMISSIONS ATTRIBUTABLE TO IMPORTS AND EXPORTS OF


ELECTRICITY AND STEAM

The consumption of power or steam (or hot water) from another company usually results in
the generation of indirect emissions – i.e., “emissions that are a consequence of activities of
the reporting company, but occur from sources owned or controlled by another company”
(WRI 2001a). Of course, virtually every raw material, energy source and service used by a
company has an indirect emissions impact. Many GHG accounting protocols, however,
selectively include indirect emissions related to electrical power and steam consumption
because they are applicable to a wide range of activities and can be a significant component
of a company’s total GHG impact. The calculation tools presented in this report, therefore,
address indirect emissions from electricity and steam (or hot water) transfers. Like most
existing protocols, these tools recommend that indirect emissions be reported separately from
direct emissions.

12.1 Emission Factors for Purchased Power and Steam

Electrical power companies and national authorities publish information on the emissions
generated in producing electrical power on a national or regional basis, so it is relatively easy

26
to estimate the indirect emissions associated with purchased power. (Sources for national
and regional information for several countries are identified in Annex D.) It is often difficult,
however, to determine whether published emission factors for electrical power include all
GHGs or only CO2 emissions. The difference is usually unimportant because CO2 represents
the great majority of the emissions in most situations. For purposes of these calculation
tools, therefore, it is assumed that purchased power emission factors address all GHGs and
are reported in CO2 equivalents. Where emission factors are available for individual gases
associated with electrical power, the individual gases can be reported separately and then
combined into carbon dioxide equivalents, or the individual emission factors can be
combined into a single CO2-equivalents emission factor.

Electrical power transmission losses vary from location to location. In some cases, they are
so significant that they provide an important rationale for distributed power generation.
However, published GHG emission factors for purchased power seldom incorporate the
effects of transmission losses. Furthermore, accepted GHG protocols seldom ask users of
electrical power to account for transmission losses. For these reasons, these tools
recommend the use of emission factors for purchased power that do not include transmission
losses. If transmission losses are particularly important, however, this can be noted in the
results and the impact can be estimated in supporting information.

In addition, some published emission factors for purchased power are “full fuel cycle”
emission factors that include upstream emissions from fuel production. Because full fuel
cycle emission factors are not the norm, these calculation tools recommend that purchased
power emission factors be based only on the emissions from the power producers and not
their upstream emissions. If companies must use full fuel cycle emission factors, to satisfy
national reporting requirements, for instance, this should be noted in the results.

12.2 Netting Imports and Exports

In some cases, facilities both import and export electricity. Because of the controversies
about estimating indirect emissions impacts, however, netting of imports and exports is a
controversial practice. The WRI/WBCSD GHG Protocol, for instance, specifically requires
that the emissions impacts of exports be reported separately, as “supporting information,”
rather than together with indirect emissions from imports, and netting is specifically
prohibited (WRI 2001a). Therefore, companies wanting to follow the WRI/WBCSD GHG
Protocol (a) should not net imports and exports, and (b) should report the emissions from
imports and exports separately. These calculation tools employ estimation methods and
reporting formats consistent with the WRI/WBCSD requirements.

12.3 Electricity Imports

To estimate indirect emissions associated with imported power, companies should use the
most appropriate purchased power emission factor available; i.e., one that reflects the
emissions generated during the production of the power being purchased. At most pulp and
paper mills, power imports are from base loads. In most cases, therefore, the base load or
average emission factor will be used rather than the marginal or peak power emission factor.
Where companies can demonstrate that a peak power emission factor (or some other

27
emission factor) is more appropriate, it can be used but the justification should be noted in
the results. In some cases, the emission factor for purchased power will reflect specific
purchasing agreements with a power supplier (e.g., for “green” power).

In cases where imported power is generated by a nearby CHP system, the emissions
associated with the imported power can be estimated using the method described in
Section 12.7. Of course, if a mill is using all of the heat and power from a CHP system, there
is no need to allocate the emissions. In such a case, if the company owns or controls the
source all the emissions will be reported as direct emissions. On the other hand, if the source
is owned or controlled by another entity all the emissions will be reported as indirect
emissions.

Example calculation: Mill purchasing electrical power.

A mill in Alberta, Canada purchases 300 TJ of electrical power (83,300 MWh) in a year’s
time. The Canadian VCR Registration Guide shows an average emission factor for
purchased power in Alberta of 0.991 kg CO2 eq./kWh (see Annex D). The indirect emissions
associated with the purchased power are estimated as follows.

· 83,300 MWh/y = 83.3 x 106 kWh/y


· (83.3 x 106 kWh/y) x (0.991 kg CO2-eq./kWh) = 82.6 x 106 kg CO2-eq./y
= 82,600 t CO2-eq./y

12.4 Electricity Exports

To be consistent with accepted GHG accounting protocols, these calculation tools suggest a
format for reporting results wherein a mill reports all direct emissions associated with the
generation of power and steam, whether the power and steam is used internally or exported.
The result tables, however, allow the mill to separately show those emissions attributable to
exported power and steam. The tables also allow companies to compare the carbon intensity
of exported power (in kg CO2/MWh) to the carbon intensity of the power on the grid into
which the power is exported. Companies may find this a helpful tool for highlighting the
beneficial environmental attributes of exports of biomass-based power and power produced
by CHP systems. Companies wanting to conform to the WRI/WBCSD GHG Protocol should
not net imports and exports or associated emissions.

These calculation tools use the method recommended by WRI/WBCSD for estimating the
emissions impact of exported power. This method involves estimating the emissions
generated by the mill to produce the exported power. Because exported electricity from mills
is usually generated in combined heat and power (CHP) systems, companies will often need
to use the methods for CHP systems (described in Section 12.7) to estimate the emissions
attributable to the exported power.

In the inventory results, companies can show the carbon intensity of the exported power or
steam (e.g., in kg CO2/MWh or kg CO2/GJ) compared to the carbon intensity of the grid into
which the power or steam is exported. To estimate the carbon intensity of the grid, the mill
should use the most appropriate grid emission factor available – i.e., one that reflects the

28
emissions assumed to be displaced by the power being exported. Because mills usually
export power into base loads (i.e., mills do not usually serve as suppliers of peaking power),
the base load emission factor will be used in most cases rather than the marginal or peak
power emission factor. Companies may use the peak or marginal emission factors, however,
if it is more appropriate.

12.5 Steam Imports

In many cases where steam is imported by a mill, the steam is produced by a nearby CHP
system. In these cases, the indirect emissions reported by the mill can be estimated using the
allocation method described in Section 12.7. In other cases, the contractual arrangement
between the mill and power producer may define how the emissions from the power plant are
to be allocated between the power and the steam sold by the power plant. In these cases, the
allocation should be explained in the results. If the imported steam is not generated in a CHP
system, best professional judgment must be used to estimate the emissions reported by the
mill. In these calculations, the heat delivered to the mill can be adjusted to reflect the amount
of heat in returned condensates. The method used to estimate the indirect emissions
associated with imported steam should be described in the results of the inventory.

12.6 Steam and Hot Water Exports

As in the case of electricity, the total on-site emissions from company-owned boilers are
shown as direct emissions whether steam or hot water is exported or not, but the emissions
associated with exported steam or hot water can be shown separately. The method for
estimating these emissions is analogous to the method used for exported electricity. The
method used to develop the estimate will depend on whether a CHP system is involved. If
steam from a boiler is exported directly without first being used in a CHP system, the
emissions from the boiler can usually be allocated in direct proportion to the amount of steam
exported (as a fraction of the total amount of steam generated by the boiler). If, however, a
CHP system is involved, the method described below should be used to allocate emissions.
In either case, the heat delivered by the mill can be adjusted to reflect the amount of heat in
returned condensates. A variety of situations will require the use of best professional
judgment. Exports of hot water are treated the same as exports of steam, on an energy
content basis (i.e., one GJ of hot water energy is assumed to be equivalent to one GJ of steam
energy, thermal losses during generation of hot water from mill-generated steam are assumed
to be negligible).

12.7 Allocating Emissions from Combined Heat and Power (CHP) Systems

Where electricity is produced by combined heat and power (CHP) systems, it may be
necessary to allocate the emissions from the CHP system to the various output energy
streams. Of course, if the mill owns the CHP system and uses all of its output, allocation is
not necessary because all of the emissions are direct emissions for the mill. In many cases,
however, a mill may either receive CHP energy from an outside provider or export a portion
of its own CHP output. For instance, if a mill is importing steam from a nearby power plant,
it is necessary to estimate the indirect emissions associated with the imported steam.
Likewise, if a mill is exporting power from a CHP system but using the steam internally, one

29
must estimate how much of the mill’s emissions are attributable to the exported electricity.
Exports of hot water are treated the same as exports of steam.

Although there are several methods for allocating emissions from CHP systems, the
“efficiency” method is recommended in these calculation tools. (Additional information on a
number of different methods is presented in Annex E.) The efficiency method is
recommended because it attempts to relate the energy outputs to the amounts of fuel used to
generate them and, by extension, to the GHGs produced in generating them. Where a
company uses an alternative method, the method should be explained in the results.

The efficiency method is one of two methods recommended by WRI/WBCSD (termed “CHP
option 1” by WRI/WBCSD) (WRI 2001b, c). There are two versions of the efficiency
method. The first, the simplified efficiency method, is less complex but involves several
assumptions about equipment efficiencies. It is expected that the simplified method will be
adequate for many mills and, therefore, it is included in this report as a default method. The
second, the detailed efficiency method, is more complicated but can use site-specific design
and operating data that companies sometimes have for CHP systems. The detailed efficiency
method is described in Annex E.

Where a mill or company has more than one CHP system, it need not allocate the emissions
from all systems using the same efficiencies for power and steam generation if there is a
basis for using different efficiencies on different CHP systems.

12.7.1 Simplified Efficiency Method

The efficiency method requires use of assumed efficiency factors for the production of power
and steam, or actual efficiency factors for each steam or power generation device based on
detailed process design and operating information. It is assumed that the efficiency of
producing hot water is the same as the efficiency of producing steam. The simplest approach
to applying the efficiency method is to assign a single efficiency factor to all power output
and a single efficiency factor to all heat (steam and hot water) output. This information is
used to compute an efficiency ratio equal to the heat production efficiency divided by the
power production efficiency. For example, if the CHP system produces steam at 80%
efficiency and power at 40% efficiency the ratio would be 2. The efficiency ratio is used
rather than the individual efficiencies because (a) it is the ratio that controls the allocation of
emissions rather than the individual efficiencies, and (b) the individual efficiencies are
constrained by the energy balance so it is not possible to specify both independently.
Emissions from the CHP system are allocated between the heat and power outputs, based on
this ratio of efficiencies, using Equations 1 and 2. This approach is referred to in this report
as the simplified efficiency method. The simplified efficiency method is the method
recommended for mills that lack, or choose not to use, detailed design and operating data
from CHP systems.

ìï H üï eH
EH = í ý ´ ET ; Reff = (Eq. 1)
ïî H + P ´ Reff ïþ eP

30
where: EH = emissions share attributable to heat production, t GHG/y
ET = total emissions from the CHP plant, t GHG/y
H = heat output, GJ/y
P = power output, GJ/y
Reff = ratio of heat production efficiency to power production efficiency
eH = assumed efficiency of typical heat production (default = 0.8)
eP = assumed efficiency of typical electric power production (default = 0.35)

The emission share attributable to electric power production is assigned from the relation:

E P = ET - E H (Eq. 2)

where: EP = emissions share attributable to electric power production

In these calculations, the heat in steam can be corrected to reflect the amount of heat in
returned condensates.

In using the simplified efficiency method, WRI/WBCSD recommends that an efficiency of


0.35 be used for power generation and 0.8 be used for steam (or hot water) generation (WRI
2001b, c), corresponding to a ratio of efficiencies (Reff) of 2.3. The first example calculation
below makes use of these recommended default efficiency factors.

Example Calculation: Allocating CHP emissions to three output streams – Simplified


efficiency method with WRI/WBCSD recommended default efficiency factors

A mill has the CHP system shown in the following figure, but it is lacking (or chooses not to
use) detailed energy balance information. Instead, the company chooses to use the simplified
efficiency method and the default efficiencies recommended by WRI/WBCSD; 0.35 for
power generation and 0.8 for steam generation (WRI 2001b, 2001c).

Fuel2
974 m3/hr nat. gas Steam P2=3 MW
HRSG
Stm. Turbine
Hs= 19.21 MW
Heat effB

H1=15 MW
Gas-fired Turbine 19,500 kg/hr steam
P1=5 MW 170°C, 7 bar
Fuel1
1538 m3/hr nat. gas

Using these assumed efficiencies, emissions can be allocated among the three outputs of the
CHP system as follows (using a basis of one hour of operation):

31
Total system emissions:

Fuel1:
CO2 (1538 m3/hr) x (0.039 GJ/m3) x (55.9 kg CO2/GJ) = 3353 kg CO2/hr
CH4 (1538 m3/hr) x (0.039 GJ/m3) x (0.0006 kg CH4 /GJ) x (21 CO2-eq. / CH4)
= 0.76 kg CO2-eq./hr
N2O (1538 m3/hr) x (0.039 GJ/m3) x (0.0001 kg N2O /GJ) x (310 CO2-eq. / N2O)
= 1.86 kg CO2-eq/hr
Total Fuel1 emissions = 3356 kg CO2-eq./hr

Fuel2:
CO2 (974 m3/hr) x (0.039 GJ/m3) x (55.9 kg CO2/GJ) = 2123 kg CO2/hr
CH4 (974 m3/hr) x (0.039 GJ/m3) x (0.0014 kg CH4 /GJ) x (21 CO2-eq. / CH4)
= 1.12 kg CO2-eq./hr
N2O (974 m3/hr) x (0.039 GJ/m3) x (0.0001 kg N2O /GJ) x (310 CO2-eq. / N2O)
= 1.18 kg CO2-eq./hr
Total Fuel2 emissions = 2126 kg CO2-eq./hr

Total CHP system emissions = 3356 + 2126 = 5482 kg CO2-eq./hr

Total system power output = P1 + P2 = 8 MW

0 .8
Reff = = 2.3
0.35

ì 15MW ü
EH = í ý ´ 5482 kgCO 2 eq = 2462 kgCO 2 eq = 20,681 t CO2 eq/y at 350 d/y operation
î 15MW + (8 MW ´ 2.3) þ

E P = 5482 kgCO eq - 2452 kgCO eq = 3020 kgCO eq = 25,368 t CO2 eq/yr at 350 d/y operation
2 2 2

Using the simplified efficiency method with default power and steam efficiency factors,
therefore, the emissions from the CHP system are allocated to the output streams in the
following percentages:
· Percentage of CHP emissions to heat output = 100*2462/5482 = 44.9%
· Percentage of CHP emissions to power output = 100*3020/5482 = 55.1%

These percentages can be used to allocate all GHG emissions from the CHP system.
Emission factors can be developed for the energy outputs:
· Emission factor for CHP heat output = (2462 kg CO2-eq./hr)/15 MW
= 164.1 kg CO2-eq./MWh
· Emission factor for CHP power output = (3020 kg CO2-eq./hr)/8 MW
= 377.5 kg CO2-eq./MWh

32
13.0 GREENHOUSE GAS EMISSIONS FROM VEHICLES AND
MISCELLANEOUS FOSSIL FUEL-FIRED EQUIPMENT

Companies often own vehicles to transport raw materials, products, wastes, and employees.
Companies may also own off-road vehicles and other types of fossil fuel-fired equipment.
The WRI/WBCSD GHG Protocol includes these emissions in its Scope 1 level of reporting
and recommends that they be included in all inventories. Because companies may want to
include these emissions in corporate GHG inventories, they are addressed in these calculation
tools. Companies should indicate in the results of the inventory whether these emissions
have been included.

13.1 Greenhouse Gases from On-Road Vehicles

Companies wanting to include these emissions can base them on either fuel consumption
statistics or information on distances traveled. Emission factors from a variety of sources
allow estimates of CO2, CH4, and N2O emissions from transportation vehicles (Annex F).
The emission factors in Annex F clearly demonstrate, however, that CH4 and N2O emissions
are a small fraction of CO2 emissions from on-road transportation sources. CO2 emission
factors for liquid transportation fuels are usually close to 70 kg CO2/GJ while, even on a
CO2-equivalents basis, combined CH4 and N2O emission factors are approximately
0.6 kg/GJ, less than 1% of the CO2 emissions. The WRI/WBCSD GHG Protocol addresses
only CO2 emissions from these sources and, given the small contribution of non-CO2 gases,
this is also the default approach suggested in these calculation tools. Where it is easier,
however, companies may want to use emission factors that include all GHGs in units of CO2-
equivalents.

If companies use fuel consumption statistics to estimate CO2 emissions, the estimates are
derived using exactly the same approach as used for stationary fossil fuel combustion
sources, described in Section 8. If, on the other hand, the company finds it more convenient
to develop the emissions estimate from statistics on distance traveled, the emission factors in
Table 7 can be used (WRI 2001d).

13.2 Greenhouse Gas Emissions from Off-Road Vehicles and Equipment

Companies may own off-road vehicles and other fossil fuel-powered equipment that they
want to include in the operational boundaries of the inventory. These sources might include
everything from forklifts to chain saws.

Fuel consumption statistics can be used to estimate CO2 emissions from these sources using
the emission factors in Table 2. Unlike on-road vehicles, however, CH4 and N2O emissions
can be a notable fraction of the GHG emissions from some of these sources. N2O emissions
for some of these sources are reported to be near 30 g/GJ or 9 kg CO2-equivalents/GJ, which
can amount to more than 10% of the CO2 emissions from such sources.

33
Table 7. Default Fuel Economy and Emission Factors for Different Types
of Mobile Sources and Activity Data (WRI 2001d)
Vehicle Characteristics Emission Factor
Vehicle Type Liters/100km Mpg gram CO2 / km
New small gas/electric hybrid 4.2 56 100.1
Small gas auto, hghwy 7.3 32 175.1
Small gas auto, city 9.0 26 215.5
Med gas auto, hghwy 7.8 30 186.8
Med gas auto, city 10.7 22 254.7
Large gas automobile, hwy 9.4 25 224.1
Large gas automobile, city 13.1 18 311.3
Med station wagon, hwy 8.7 27 207.5
Med station wagon, city 11.8 20 280.1
Mini van, hwy 9.8 24 233.5
Mini van, city 13.1 18 311.3
Large van, hwy 13.1 18 311.3
Large van, city 16.8 14 400.2
Mid size pick-up truck, hwy 10.7 22 254.7
Pick-up truck, city 13.8 17 329.6
Large pick-up truck, hwy 13.1 18 311.3
Large pick-up truck, city 15.7 15 373.5
LPG automobile 11.2 21 266
Diesel automobile 9.8 24 233
Gasoline light truck 16.8 14 400
Gasoline heavy truck 39.2 6 924
Diesel light truck 15.7 15 374
Diesel heavy truck 33.6 7 870
Light motorcycle 3.9 60 93
Diesel bus 35.1 6.7 1035

Emission factors for these sources have been published in a number of places. Some of the
available information is summarized in Annex F. IPCC’s guidelines contain several different
sets of emission factors without recommending a single set (IPCC 1997c). Table 8 is from
one of the sources cited by IPCC. The CO2 factors in the table are from Table 2 and are
slightly different than those in the original table in IPPC 1997c. The table also includes
overall CO2-equivalent emission factors developed using the IPCC global warming potentials
for CH4 (21) and N2O (310). Emission factors for CH4 and N2O from these sources vary
from one protocol to another. The differences in N2O, in particular, can impact the CO2-
equivalents by as much as 10%. Where companies need precise estimates for these sources,
it is recommended that the various sources discussed in Annex F be examined to determine
which emission factors are most appropriate. In most cases, however, the emission factors in
Table 8 will be adequate.

34
Table 8. Fuel Consumption-Based Emission Factors
for Non-Road Mobile Sources and Machinery (IPCC 1997c)
(IPCC Revised 1996 Guidelines taken from EMEP/CORINAIR)
Source and Engine Type CO2 kg/TJ CH4 kg/TJ N2O kg/TJ CO2-equiv. kg/TJ
Forestry – diesel 73,400* 4 30 82,800
Industry – diesel 73,400* 4 30 82,800
Railways – diesel 73,400* 4 30 82,800
Inland waterway – diesel 73,400* 4 30 82,800
Marine – diesel 73,400* 7 2 74,200
Industry – gasoline 68,600* 50 2 70,300
4-stroke
Forestry – gasoline 68,600* 170 0.4 72,300
2-stroke
Industry – gasoline 68,600* 130 0.4 71,500
2-stroke
Inland waterway – 68,600* 40 2 70,100
gasoline 4-stroke
Inland waterway – 68,600* 110 0.4 71,000
gasoline 2-stroke
* from Table 2, corrected for unburned carbon

Example calculation: GHG emissions from on-site vehicles and equipment.

Based on purchasing records, a mill estimates the amounts of fuel purchased over a year to
fuel on-site vehicles and equipment. It applies the largest of the emission factors shown in
Table 1 and estimates that the emissions are much less than 0.5% of the mill’s emissions.
Rather than trying to develop a more accurate estimate, the mill decides to report in the
results only that the emissions from this source are non-material because they represent less
than 0.5% of the total emissions.

35
Example calculation: GHG emissions from a company’s forestry operations and wood
transport fleet.

A company’s fuel purchasing records indicate that, in a year, the following amounts of fuel
are consumed by the company’s woodlands operations and its fleet of trucks used to transport
wood to the mill:
· Gasoline = 10,000 l – The company estimates that approximately 90% of this is used in
4-stroke engines and 10% is used in 2-stroke engines in forestry equipment.
· Diesel = 200,000 l
The heat content of the gasoline is estimated to be 0.034 GJ/l and the heat content of the
diesel fuel is 0.038 GJ/l.
The company decides to use the CO2-equivalent emission factors in Table 8 to estimate
emissions.

· Gasoline used in 4-stroke engines = 10,000 l/y x 0.9 = 9,000 l/y


· 9,000 l/y x 0.034 GJ/liter = 306 GJ/y = 0.306 TJ/y
· 0.306 TJ/y x 70,300 kg CO2-equiv./TJ = 21,500 kg CO2-equiv./y = 21.5 t CO2-equiv./y

· Gasoline used in 2-stroke engines = 10,000 l/y x 0.1 = 1,000 l/y


· 1,000 l/y x 0.034 GJ/liter = 34 GJ/y = 0.034 TJ/y
· 0.034 TJ/y x 72,300 kg CO2-equiv./TJ = 2460 kg CO2-equiv/y = 2.5 t CO2-equiv./y

· Diesel used = 200,000 l/y


· 200,000 l/y * 0.038 GJ/l = 7600 GJ/y = 7.6 TJ/y
· 7.6 TJ/y x 82,800 kg CO2-equiv./TJ = 629,000 kg CO2-equiv./y = 629 t CO2-equiv./y

Total GHG emissions from company-owned forestry operations and wood trucks
· 21.5 + 2.5 + 629 = 653 t CO2-equiv./y

14.0 GREENHOUSE GAS EMISSIONS FROM MILL WASTE IN LANDFILLS

These calculation tools have been developed assuming that many companies will include
mill-owned landfills within the inventory boundaries. These tools can also be used in cases
where a mill’s process waste is being disposed in a municipal solid waste landfill and the
company is interested in estimating the mill’s contribution to the municipal landfill
emissions. The reporting format, however, has been prepared assuming that only the
emissions from company-owned landfills will be reported.

As is the case with most widely accepted protocols, only CH4 emissions are addressed in
these tools because CO2 from landfills is composed of biomass carbon (not counted as a
greenhouse gas) and N2O emissions are assumed to be negligible.

An emission factor for landfilled waste was presented in Table 1. This factor is based on a
number of conservative assumptions and, in most cases, is expected to produce estimates that
are higher than the actual emissions attributable to landfilled mill waste. The emission factor
can be useful in deciding whether landfill emissions are material to the results of the

36
inventory. The methods described below, however, are recommended for preparing an
estimate to use in the inventory results.

14.1 Using Data from Landfill Gas Collection Systems

In some cases, company landfills are capped with low permeability cover material and the
landfill gas is collected. In many of these situations, the amounts of methane collected and
destroyed can be estimated from site-specific data. IPCC’s recommended approach uses this
information only indirectly. IPCC recommends that companies estimate landfill gas
emissions by first estimating total gas generation (using one of several mathematical models
discussed below) and then subtracting the amounts of methane captured and burned. The
difference between the two is assumed to be emitted. The problem with this approach is that,
because of the large uncertainties in estimating methane generation, the amounts burned
could easily be greater than the amounts that the company estimates were generated,
resulting in a negative release. It is equally possible that the comparison of estimated
generation rates to measured collection rates could suggest impossibly low collection
efficiencies, due only to the uncertainties in estimating methane generation.

An alternative approach is available to companies that measure the amounts of methane


captured in efficient collection systems. The alternative approach is to estimate the
collection efficiency of the system and then back calculate the amounts of methane
generated. For instance, if a mill with a capped landfill has determined that its collection
system collects 90 tons of methane per year and the mill estimates that the collection
efficiency is 90%, it means that 100 tons of methane were generated.

The problem with this approach is that the effectiveness of landfill gas collection systems is
variable and uncertain. Reported collection efficiencies range from 60 to 85% (USEPA
1998). This variability and uncertainty has caused IPPC to take the position that “the use of
undocumented estimates of landfill gas recovery potential is not appropriate, as such
estimates tend to overestimate the amount of recovery” (IPCC 2000a). Nonetheless, this
approach is built around a measured value – the amount of gas collected. For this reason, it
is reasonable to expect that in some, if not many, cases it will yield more accurate estimates
than IPCC’s default methodology. This is especially true for mill landfills because of the
limited data available for deriving the parameter values needed to use IPCC’s mathematical
models for estimating emissions.

Therefore, in these calculation tools it is recommended that where mill landfills are covered
with low permeability caps and equipped with landfill gas collection systems that are
constructed and operated to normal standards, the methane generation rates should be back
calculated from (a) measurements of the amounts of methane collected; and (b) measured or
assumed collection efficiency. A default collection efficiency of 75% has been used by some
authorities and is recommended here unless site-specific collection efficiency data are
available (USEPA 1998). These calculation tools also assume that all of the methane that is
captured and burned is converted to biomass CO2 and therefore does not have to be included
in the inventory.

37
Using these default values and assumptions, estimates of methane releases can be developed
using Equation 3.

CH4 (m3/y) released to the atmosphere =


[(REC / FRCOLL) * (1 – FRCOLL) * FRMETH * ( 1 – OX )] +
[ REC * FRMETH * ( 1 – FRBURN)] (Eq. 3)

where: REC = amount of landfill gas collected, determined on a site-specific basis, m3/y
FRCOLL = fraction of generated landfill gas that is collected, default is 0.75
FRMETH = fraction of methane in landfill gas, default is 0.5
OX = fraction of methane oxidized in the surface layer of the landfill, default is 0.1
FRBURN = fraction of collected methane that is burned, site-specific determination

14.2 Estimating Landfill Methane Emissions at Mill Landfills without Gas Collection
Data

14.2.1 Simplified First Order Decay Approach

Where the approach described above cannot be used, it is recommended that companies
employ the first order decay model approach for estimating landfill gas emissions using
parameter values derived for pulp and paper mill landfills. This approach is the default
method recommended by IPCC and is used by a number of national authorities (IPCC
2000a). It can be used to estimate CH4 emissions from active and inactive landfills.

In cases where the annual deposits are constant (or are assumed to be) IPCC’s default method
reduces to two equations. This simplified approach should be adequate unless the amounts or
types of waste being landfilled have changed significantly from year to year (e.g., a new
deinking mill is built on-site) or the landfill design or operation has been changed in a way
that would significantly impact methane generation or release (e.g., a gas collection system is
installed). The simplified approach is as follows.

CH4 (m3/y) generated from all waste in the landfill = R L0 (e-kC - e-kT) (Eq. 4)

where: R = average amount of waste sent to landfill per year, Mg/y


L0 = ultimate methane generation potential, m3/Mg waste
k = methane generation rate constant, 1/y
C = time since landfill stopped receiving waste, y
T = years since landfill opened, y
(Note: R and L0 can be in units of wet weight, dry weight, degradable organic
carbon, or other units but the units for R and L0 must be the same.)

Not all methane that is generated is subsequently released to the atmosphere. To estimate
atmospheric releases, use the result from Equation 4 in Equation 5. For landfills with
modern gas collection and combustion systems but no measurements of quantities of gas
collected, the amount of methane recovered can be assumed to be 75% of that generated
(USEPA 1998).

38
CH4 (m3/y) released =
[(CH4 generated–CH4 recovered)*(1–OX)]+[CH4 recovered*(1-FRBURN)] (Eq. 5)

where: CH4 generated = from Equation 4


CH4 recovered = amount of methane collected, site-specific determination
OX = fraction oxidized in the surface layer of the landfill before escaping, usually
assumed to be 0.1
FRBURN = fraction of collected methane that is burned, site-specific determination

If the amounts being landfilled have changed significantly or if the landfill design has been
altered so that some of the parameter values would have changed substantially over time, a
more involved approach may be needed. To deal with these more complicated situations,
many protocols recommend modeling the gas generated annually from each year’s deposits
and summing the amounts that are predicted to occur in the current year. This more detailed
analysis is described in Section 14.2.2.

Annex G identifies a number of sources for the parameter values L0 and k needed in
Equation 4. Unfortunately, the values vary considerably from one protocol to the next and
the values are based on very few data. For situations where pulp and paper mill wastewater
treatment sludge is a major constituent of the waste, reasonable values for the rate constant,
k, fall in the range of 0.01/yr to 0.1/yr while those for L0 fall between 50 and 200 m3/Mg.
Research is currently underway in the US that should help narrow these ranges. Initial
indications are that the amounts of gas generated in mill landfills are less than would be
predicted using parameter values developed for municipal solid waste (NCASI 1999). With
this knowledge, it is recommended that until the current research is completed, and unless
companies have country-specific or site-specific factors that are more appropriate for their
mill wastes, companies should use the parameter values shown in Table 9.

Table 9. Recommended Default Values for k and L0 for


Estimating Mill Landfill Methane Emissions
Parameter Default Value
k 0.03 y-1
L0 100 m3/Mg dry weight of waste

14.2.2 Detailed First Order Decay Approach

To allow year-to-year variations in the amounts of waste sent to landfill, IPCC suggests a
variation of this approach. Using this variation, starting in year one, calculate how much
methane will be generated in each subsequent year by waste deposited in that year using
Equations 6 and 7.

CH4 generated in a given year by waste deposited in an earlier year (m3/y)


= k Ry L0 (e-k[T-Y]) (Eq. 6)

where k = methane generation rate constant, 1/yr


RY,= the amount of waste sent to landfill in year Y, Mg/yr

39
L0,= ultimate methane generation potential, m3/Mg waste
T = year for which emissions are being estimated, given in terms of years since the
landfill opened
Y = year after landfill opened that waste was disposed
Thus (T-Y) is equal to the number of years the waste has been in place prior to the
year for which emissions are being estimated.

CH4 (m3/y) released = [(CH4 generated–CH4 recovered)*(1–OX)]


+[CH4 recovered*(1-FRBURN)] (Eq. 7)

where: CH4 generated = from Equation 5


CH4 recovered = amount of methane collected, site-specific determination
OX = fraction oxidized in the surface layer of the landfill before escaping, usually
assumed to be 0.1
FRBURN = fraction of collected methane that is burned, site-specific determination

The calculations are performed by estimating how much waste was deposited every year
since the landfill was opened. IPCC indicates that for very old landfills, it is possible to limit
the retrospective period to one starting at least three waste degradation half-lives before the
current year. Given the slow degradation observed in many mill sludges, 25 years is
probably the minimum that would satisfy this criterion. For each year’s deposit, the amount
of methane released that year and each following year is estimated. In subsequent years, the
amount of methane released is the sum of the amounts estimated from each prior year’s
deposits that were projected to occur in that year.

In year 1, amount A is deposited and it is estimated that in years 1, 2, 3, … it will release X1,
X2, X3, … tons of methane, respectively. The reported emissions for year 1 are X1 tons of
methane. In year 2, amount B is deposited and it is estimated that in years 2, 3, 4, … it will
release Y2, Y3, Y4, …tons of methane, respectively. The emissions reported for year 2 are X2
plus Y2 tons methane. In year 3, amount C is deposited and it is estimated that in years 3, 4,
5, … it will release Z3, Z4, Z5, … tons of methane, respectively. The reported emissions for
year 3 are X3 plus Y3 plus Z3 tons of methane. This process repeats itself every year.

The values for k and L0 are the same as those used in the simplified first order approach, as
presented in Table 9.

40
Example Calculation: Emissions from a mill landfill with a modern low-permeability cap
and gas collection system. The collected gas is burned.

Measurements have been made on a landfill gas collection system. The system is collecting
820,000 standard m3/y and the gas is 47% methane by volume. The mill has no site-specific
data on the efficiency of the gas collection system, so it uses the recommended default value
of 75%. It also uses the default assumption that 10% of the uncollected gas is oxidized
before escaping to the atmosphere.

· methane collected = 820,000 m3/y x 0.47 = 385,000 m3/y


· methane generated = (385,000 m3/y) / 0.75 = 513,000 m3/y
· methane released = (513,000 – 385,000)m3/y x (1- 0.1) = 115,000 m3/y = 115 x 106 l/y
· methane released = (115 x 106 l/y) / 22.4 l/g-mole = 5.13 x 106 g-mole/y
· methane released = (5.13 x 106 g-mole/y) x 16 g/g-mole = 82 x 106 g/y = 82 t CH4/y
Using the IPCC GWP (21), this is equal to 1720 t CO2-equiv./y

Example calculation: Emissions from 20 year old landfill receiving mill wastewater
treatment solids and ash. The landfill does not have a gas collection system.

A mill landfills 50 ton per day of solid waste composed primarily of wastewater treatment
plant solids, ash, and other miscellaneous waste typical of a kraft mill. The mill generates
waste 350 days a year. The landfill has been in use for 20 years and is still active. The
landfill does not have a gas recovery system. The mill uses the default values for k and L0
shown in Table 9 (100 m3/Mg for L0 and 0.03 y-1 for k).

R = 50 Mg/d x 350 d/y = 17500 Mg/y


L0 = 100 m3/Mg
k = 0.03/y
C=0y
T = 20 y

· methane generated (m3/y) = 17,500 x 100 x (e-0.03 x 0 – e-0.03 x 20) = 790,000 m3/y
· density of methane (0oC and 1 atm. pressure) = 0.72 kg/m3 (from Perry’s Chemical
Engineers’ Handbook)
· methane generated (kg/y) = 790,000 m3/y x 0.72 kg/m3 = 568,000 kg/y = 568 t/y
· assume 10% oxidation in landfill cover
· methane released = 568 t/y x (1 – 0.1) = 511 t CH4 /y released
Using the IPCC GWP (21), this equals 10,700 t CO2-equiv./y

Note that the Table 1 emission factor would have yielded an estimate of 50 t/d x 350 d/y x
3,500 kg/t = 61,250,000 kg/y = 61,250 t CO2-equiv./y, over five times the estimate developed
using the more refined approach.

41
15.0 GREENHOUSE GAS EMISSIONS FROM ANAEROBIC TREATMENT OF
WASTEWATER OR SLUDGE

Most existing GHG protocols address waste treatment plant emissions only from anaerobic
treatment and digestion processes. Therefore, these calculation tools have been developed
assuming that emissions from other types of wastewater and sludge treatment processes are
negligible. Although aerobic and facultative treatment systems may have zones with
depleted dissolved oxygen, methane generation rates in aerated stabilization basins, activated
sludge systems, and their associated retention ponds would be expected to be much lower
than those estimated for anaerobic systems. In any event, due to lack of data current GHG
protocols to not attempt to estimate emissions from aerobic and facultative treatment
operations.

Even for anaerobic systems, only CH4 emissions need to be estimated. The CH4 emissions
from mill-owned anaerobic systems will be reported as direct emissions. The CO2 emitted
from wastewater and sludge treatment operations contains biomass carbon which is not
included in most GHG inventories. In addition, N2O emissions from treatment plants have
been found to be small, and probably occur only after the wastewater is discharged (IPCC
1997c).

15.1 Anaerobic Treatment Operations where Off-Gases are Captured

In many cases, anaerobic treatment systems are covered and the gases are collected and
burned. One of the purposes of these collection systems is the prevention of odors, and to
accomplish this objective the systems must be highly efficient. For purposes of a GHG
inventory it is reasonable to assume, therefore, that where methane emissions from anaerobic
treatment operations are captured and burned, the collection and destruction is complete and
no methane is emitted. Because the CO2 produced in burning the CH4 contains biomass
carbon, it does not need to be reported in the inventory results. If circumstances at a mill
suggest that non-trivial amounts of methane are escaping collection, the mill may need to
undertake efforts to account for these releases, but such circumstances are expected to be
unusual at mills that collect and burn these gases.

Of course, if the gases are collected but released to the atmosphere rather than being burned,
they should be included in the inventory.

15.2 Anaerobic Treatment Operations where Off-Gases are Released to the


Atmosphere

Where off-gases from anaerobic treatment operations are not collected and burned, it is
necessary to estimate the releases of methane to the atmosphere. In some cases, for instance
where the gases are released through a vent in a covered vessel, the releases can be measured
directly. In most other cases, they must be estimated.

These calculation tools suggest the use of the IPCC default methodology described in the
May 2000 Good Practices document and shown in Equation 8 (IPCC 2000a). Although the
IPCC document allows the equation to be applied to systems that are not completely
anaerobic (by multiplying the result by an arbitrary adjustment factor of less than one), there

42
are no data currently available to support the selection of the adjustment factor. It is
recommended, therefore, that methane emissions only be estimated from anaerobic treatment
or sludge digestion systems until factors for other types of systems are available.

Anaerobic Treatment Plant Methane Emissions (kg/y) = ( OC x EF ) – B (Eq. 8)

where: OC = BOD or COD of the feed to the anaerobic system, kg/year


EF = emission factor, default values = 0.25 kg CH4/kg COD in the feed or 0.6 kg
CH4/kg BOD in the feed (or another BOD-based factor developed by multiplying
the COD-based factor of 0.25 kg CH4/kg COD by the site-specific COD/BOD ratio)
B = methane captured and burned, kg CH4/year, determined on a site-specific basis

If the solids are handled separately, emissions from sludge digestion would be calculated
using Equation 9. In cases where sludge is burned, it is included in the calculations for GHG
emissions from biomass burning discussed in Section 11.

Anaerobic Sludge Digestion Plant Methane Emissions (kg/y) = (OCs x EFs) - B (Eq. 9)

where: OCs = organic content of the sludge


EFs = emission factor, in units consistent with OCs - IPCC’s default value is
0.25 kg CH4/kg COD in the sludge feed
B = methane captured and burned, kg CH4/yr, determined on a site-specific basis

Example calculation: Recycled paperboard mill with anaerobic treatment but no gas
recovery.

A recycled paperboard mill uses an anaerobic treatment plant to treat wastewater containing
10,000 kg COD/d. The mill generates wastewater 300 days per year. The IPCC default
value for CH4 generation from anaerobic treatment systems is 0.25 kg CH4/kg COD in the
feed. The CH4 emissions are calculated as follows.

· OC = 10,000 kg/d x 300 d/y = 3,000,000 kg COD/y


· CH4 generated = 3,000,000 kg COD/y x 0.25 kg CH4/kg COD = 750,000 kg CH4/y
= 750 t CH4/y
Using the IPCC GWP (21), this equals 15,750 t CO2-equiv./y

16.0 PRESENTING THE RESULTS OF THE INVENTORY

These calculation tools include a suggested format for summarizing inventory results.
Companies may find other formats more convenient or appropriate, however, and are free to
choose the format best suited to their needs. It is important that the output from using these
calculation tools be (a) disaggregated to the extent possible to ensure transparency; and
(b) accompanied with key information needed to interpret the results.

Four tables that companies may use to present the results of the inventory are presented in the
following pages. Table 10 provides a format for identifying the core operations that have
been included within the operational inventory boundaries and who owns them. Companies

43
are encouraged to include any additional information that will help explain the boundaries or
the results of the inventory.

Table 11 is for recording direct emissions. These are emissions within the boundaries of the
inventory from sources owned or controlled by the company. The company is free to select a
method to determine ownership of emissions from sources only partly owned or controlled
by the company, but the method should be described in the presentation of results.

Table 12 is for recording indirect emissions (i.e., emissions from sources within the
operational boundaries of the inventory but owned by another entity), emissions attributable
to exports of power and steam, and imports and exports of fossil fuel-derived CO2.
Companies are encouraged to use this form to characterize the impact of electricity and steam
transfers as well as outsourced operations (power islands in particular) that have a significant
impact on a facility’s greenhouse gas profile.

Table 13 is for recording the emission factors used to prepare the inventory. Companies are
encouraged to include this information to make the results of the inventory more transparent.

Tables 14 through 17 illustrate the use of the reporting tables on an example mill inventory.
The schematic presented in Figure 1 illustrates the various sources and categories of
emissions which may be included in an emissions inventory.

Companies wanting to prepare reports that meet the requirements of the WRI/WBCSD GHG
Protocol will also need to report releases of climate-neutral CO2 from biomass combustion as
supporting information. Annex I includes a table (Table 4) that can be used for this purpose.

An Excelâ workbook that performs the calculations described in this report is available. The
completed workbook represents another way to convey the results of the inventory.

44
Table 10. Table to Report Operational Boundaries of the Inventory
This matrix may be used to indicate which core operations are included within the boundaries of the inventory
and their ownership. Provide a general description of the boundaries, any additional information needed to
explain them and then put an “X” in appropriate boxes.
General description of Operational Boundaries:

Use this space to provide additional information helpful to understanding the

Completely Owned

Partly Owned by

Owned by other
operational boundaries of the inventory, including the method used to allocate

by Company
emissions from partly-owned or –controlled sources. Attach additional pages if needed.

Company

entity
Harvesting
Wood/chip/bark/wastepaper/other raw material transportation vehicles
Product, by-product or waste transportation vehicles
Debarking
Chipping
Mechanical pulping
Chemical pulping – kraft
Chemical pulping – sulfite
Chemical pulping – other
Semichemical pulping
Recovery furnace – kraft
Liquor furnace – sulfite
Liquor furnace – semichem
Lime kiln or calciner
Incinerators for noncondensible gases, etc.
Wastepaper pulping and cleaning
Deinking
Bleaching of chemical or semichemical pulp
Brightening of deinked pulp
On-site preparation of chemicals used in core operations (e.g., ClO2 or O3)
Paper and/or paperboard production
Coating (including extrusion coating)
Roll trimming, roll wrapping, sheet cutting
On-site power and steam boilers
On-site combustion turbines
Gas-fired infrared dryers
Other fossil fuel-fired dryers
Wastewater treatment operations
Sludge processing
Landfill receiving mill waste
Air emissions control devices
On-road vehicles
Off-road vehicles and machinery
Normal offices/workspace for mill employees
Other Operation – describe:
Other Operation – describe:
Other Operation – describe:

45
Table 11. Table to Report GHG Inventory Results – Direct Emissions
emissions from sources that are wholly or partially owned or controlled by the company
Write “NA” to show an item is not applicable. For instance, if the mill Total Direct Emissions
has no biomass boilers, write NA in the boxes for biomass combustion. – metric tons
Where emissions have been determined to be insignificant or non- CO2 CH4 N2O CO2
material, write “NM” and explain the basis for the determination in a Equiv 1
footnote.
Process and Energy-Related Emissions
1 Stationary Fossil Fuel Combustion
2 Biomass Combustion 0.0
3 Make-up Chemicals (CaCO3 and Na2CO3)
Transportation and machinery emissions
4 On-road vehicles
5 Off-road vehicles and machinery
Waste management emissions
6 Landfill emissions from mill wastes 0.0
7 Anaerobic wastewater treatment systems 0.0
8 Other Direct Emissions not included above – Explain:

Total Direct Emissions (Sum of lines 1 through 8)


Explain the method used to determine ownership of emissions from sources not completely owned by the
company. Use the WRI/WBCSD GHG Protocol for guidance on determining ownership.

Include any other information that is needed to understand the inventory results:

1
CO2 Equivalents are calculated multiplying individual gases by IPCC GWP values, CO2=1, CH4=21,
N2O=310, and summing across all three gases. It is acceptable to use emission factors for CO2-equivalents
rather than estimating the three gases individually.

46
Table 12. Table to Report GHG Inventory Results – Indirect Emissions
emissions attributable to power/steam exports, and imports/exports of fossil CO2
Write “NA” to show an item is not applicable. Where emissions have Metric tons
been determined to be insignificant or non-material, write “NM” and CO2 CH4 N2O CO2
explain the basis for the determination in a footnote. Equiv 1

Indirect emissions related to electricity and steam imports, including those from outsourced power islands
1 Indirect Emissions related to electricity imports
2 Indirect Emissions related to steam imports
3 Total indirect emissions from power/steam imports
(Sum of lines 1 through 2)
Other Indirect Emissions
4 Description of other indirect emissions included in inventory:

Emissions associated with exported electricity and steam


5 Emissions related to electricity exports
Carbon intensity of electricity exports (kg CO2/MWh)
Carbon intensity of grid receiving electricity exports (kg CO2/MWh)
Method used to estimate GHG intensity of grid:

6 Emissions related to steam exports


7 Total emissions attributable to exports (Sum of lines 5 and 6)
Imports and Exports of fossil fuel-derived CO2
8Imports of CO2 (e.g., for neutralization)
9Exports of fossil fuel-derived CO2 (e.g., to PCC Plants)
Note 1: This includes only the fraction of CO2 exports that can be traced to fossil fuels. Exports of
biomass-derived CO2 are reported in Annex I – Supporting Information on Biomass.
Note 2: This exported CO2 should not be reported as an emission in Table 11.
Explain the method used to determine ownership of emissions from sources partly owned by the company.
Use the WRI/WBCSD GHG Protocol for guidance on determining ownership.

Include any other information needed to understand the inventory results:

1
CO2 Equivalents are calculated multiplying individual gases by IPCC GWP values, CO2=1, CH4=21,
N2O=310, and summing across all three gases. It is acceptable to use emission factors for CO2-equivalents
rather than estimating the three gases individually

47
Table 13. Table to Report Emission Factors (EF) Used to Prepare the Inventory (show units)
CO2 CH4 N2O CO2 Equiv Source of EF

Fossil Fuel Combustion


Fuel Combustion Units

Biomass Combustion
Fuel Combustion Unit
0.0
0.0
0.0
0.0
0.0
0.0
0.0
Waste Management
Landfill 1 emissions: % of Gas Collected = “k”= “L0”=
Landfill 2 emissions: % of Gas Collected = “k”= “L0”=
Landfill 3 emissions: % of Gas Collected = “k”= “L0”=
Anaerobic Treatment emissions: “EF”=
Electrical Power and Steam Imports
Emissions factors for imported electricity

Emission factors for imported steam

48
Table 14. Example GHG Inventory Results - Operational Boundaries of the Inventory
This matrix may be used to indicate which core operations are included within the boundaries of the inventory
and their ownership. Provide a general description of the boundaries, any additional information needed to
explain them and then put an “X” in appropriate boxes.
General description of Operational Boundaries:

Use this space to provide additional information helpful to understanding the

Completely Owned

Partly Owned by

Owned by other
operational boundaries of the inventory, including the method used to allocate

by Company
emissions from partly-owned or –controlled sources. Attach additional pages if needed.

Company

entity
Small amounts of purchased power for a third party waste paper sorting operation are included in the
mill’s inventory results. Also, an on-site gas turbine CHP system owned by another company supplies
the mill with power and steam, but much of the power from the system is sold. The emissions are
allocated using the simplified efficiency method. When the mill is down, we sometimes continue to
generate biomass power in mill-owned condensing turbines and sell it to the grid.

Harvesting X
Wood/chip/bark/wastepaper/other raw material transportation vehicles X
Product, by-product or waste transportation vehicles
Debarking X
Chipping X
Mechanical pulping
Chemical pulping – kraft X
Chemical pulping – sulfite
Chemical pulping – other
Semichemical pulping
Recovery furnace – kraft X
Liquor furnace – sulfite
Liquor furnace – semichem
Lime kiln or calciner
Incinerators for noncondensible gases, etc. X
Wastepaper pulping and cleaning X
Deinking X
Bleaching of chemical or semichemical pulp X
Brightening of deinked pulp X
On-site preparation of chemicals used in core operations (e.g., ClO2 or O3) X
Paper and/or paperboard production X
Coating (including extrusion coating) X
Roll trimming, roll wrapping, sheet cutting X
On-site power and steam boilers X
On-site combustion turbines X
Gas-fired infrared dryers X
Other fossil fuel-fired dryers
Wastewater treatment operations X
Sludge processing X
Landfill receiving mill waste X
Air emissions control devices X
On-road vehicles X
Off-road vehicles and machinery X
Normal offices/workspace for mill employees X
Other Operation – describe: X
On site commercial wastepaper collection and sorting operation
Other Operation – describe:
Other Operation – describe:

49
Schematic for Example Inventory Results
See Supporting
Information on Biomass
Direct Emissions
CH4
CO2, CH4, Emissions of
and N2O CO2, CH4, CO2 from fossil fuel, Climate Neutral
CH4, and N2O CO2 from Biomass
and N2O
Fossil fuel-derived
CO2 exported to
PCC Plant
Company-
owned CO2
off-site
vehicles and
equipment
Lime Kiln Biomass
that use Company-
and
fossil fuels owned on-site
CO2 from combination
boilers, vehicles
make-up fuel-fired
and
carbonates boilers
equipment
used in the
that use only
pulp mill
fossil fuels

On-site production of power and


steam in gas turbine CHP system Emissions from
owned by another company landfill

Mill
Site CO2 from fossil fuel, Off-site production of
CH4, and N2O purchased power from fossil
fuels,

CO2 from fossil fuel,


Indirect Emissions CH4, and N2O

Figure 1. Schematic for Example Inventory Results

50
Table 15. Example GHG Inventory Results - Direct Emissions
emissions from sources that are wholly or partially owned or controlled by the company
Write “NA” to show an item is not applicable. For instance, if the mill Total Direct Emissions
has no biomass boilers, write NA in the boxes for biomass combustion. – metric tons
Where emissions have been determined to be insignificant or non- CO2 CH4 N2O CO2
material, write “NM” and explain the basis for the determination in a Equiv 1
footnote.
Process and Energy-Related Emissions
1 Stationary Fossil Fuel Combustion 720000 100 80 746900
2 Biomass Combustion 0.0 120 40 14920
3 Make-up Chemicals (CaCO3 and Na2CO3) 5500 0 0 5500
Transportation and machinery emissions
4 On-road vehicles >>> >>> >>> 320
5 Off-road vehicles and machinery NM* NM* NM* NM*
Waste management emissions
6 Landfill emissions from mill wastes 0.0 511 NA 10730
7 Anaerobic wastewater treatment systems 0.0 NA NA NA
8 Other Direct Emissions not included above – Explain:

Total Direct Emissions (Sum of lines 1 through 8) >>> >>> >>> 778370
Explain the method used to determine ownership of emissions from sources not completely owned by the
company. Use the WRI/WBCSD GHG Protocol for guidance on determining ownership.

Include any other information that is needed to understand the inventory results:

* These emissions are uncertain but were estimated using the highest fuel consumption and emission factor data available
and were determined to be less than 0.5% of the mill’s emissions. They are therefore reported as non-material (NM).

1
CO2 Equivalents are calculated multiplying individual gases by IPCC GWP values, CO2=1, CH4=21,
N2O=310, and summing across all three gases. It is acceptable to use emission factors for CO2-equivalents
rather than estimating the three gases individually.

51
Table 16. Example GHG Inventory Results - Indirect Emissions
emissions attributable to power/steam exports, and imports/exports of fossil CO2
Write “NA” to show an item is not applicable. Where emissions have Metric tons
been determined to be insignificant or non-material, write “NM” and CO2 CH4 N2O CO2
explain the basis for the determination in a footnote. Equiv 1

Indirect emissions related to electricity and steam imports, including those from outsourced power islands
1 Indirect Emissions related to electricity imports >>> >>> >>> 72000
2 Indirect Emissions related to steam imports >>> >>> >>> 12400
3 Total indirect emissions from power/steam imports 84400
(Sum of lines 1 through 2)
Other Indirect Emissions
4 Description of other indirect emissions included in inventory: 0

Emissions associated with exported electricity and steam


5 Emissions related to electricity exports 0 6 2 746
Carbon intensity of electricity exports (kg CO2/MWh) <10
Carbon intensity of grid receiving electricity exports (kg CO2/MWh) 726
Method used to estimate GHG intensity of grid:

6 Emissions related to steam exports 0 0 0 0


7 Total emissions attributable to exports (Sum of lines 5 and 6) 0 6 2 746

Imports and Exports of fossil fuel-derived CO2


8Imports of CO2 (e.g., for neutralization) 0
9Exports of fossil fuel-derived CO2 (e.g., to PCC Plants) 21000
Note 1: This includes only the fraction of CO2 exports that can be traced to fossil fuels. Exports of
biomass-derived CO2 are reported in Annex I – Supporting Information on Biomass.
Note 2: This exported CO2 should not be reported as an emission in Table 11.
Explain the method used to determine ownership of emissions from sources partly owned by the company.
Use the WRI/WBCSD GHG Protocol for guidance on determining ownership.

Include any other information needed to understand the inventory results:

1
CO2 Equivalents are calculated multiplying individual gases by IPCC GWP values, CO2=1, CH4=21,
N2O=310, and summing across all three gases. It is acceptable to use emission factors for CO2-equivalents
rather than estimating the three gases individually

52
Table 17. Example GHG Inventory Results - Emission Factors (EF) Used to Prepare the Inventory
CO2 CH4 N2O CO2 Equiv Source of EF

Fossil Fuel Combustion


Fuel Combustion Units
Gasoline forestry equip. NA NA NA 74.7 t/TJ Table 10
Diesel fuel trucks and machinery NA NA NA 82.4 t/TJ Table 10
Coal boiler 94.6 0.7 1.6 NA Tables 4 & 5
t/TJ kg/TJ kg/TJ
Natural gas lime kiln 56.1 2.7 0 NA Table 6
t/TJ kg/TJ
Residual fuel oil recovery furnace 77.4 2.5 2 NA Tables 4 & 8
t/TJ kg/TJ kg/TJ

Biomass Combustion
Fuel Combustion Unit
bark & wood waste boiler 0.0 12 4 NA Table 8
kg/TJ kg/TJ
0.0
0.0
0.0
0.0
0.0
0.0
Waste Management
Landfill 1 emissions: % of Gas Collected = 75 “k”= 0.03 “L0”= 100 m3/Mg dry wt.
Landfill 2 emissions: % of Gas Collected = “k”= “L0”=
Landfill 3 emissions: % of Gas Collected = “k”= “L0”=
Anaerobic Treatment emissions: “EF”=
Electrical Power and Steam Imports
Emissions factors for imported electricity
Power purchased from local grid >>> >>> >>> 726 kg CO2 Information from
per MWh power supplier

Emission factors for imported steam

53
REFERENCES

AEA Technology. 2001. UK greenhouse gas inventory 1990-1999: Annual report for
submission under the Framework Convention on Climate Change. Oxfordshire, England:
National Environmental Technology Centre.
http://ariadne.aeat.co.uk/netcen/airqual/reports/ghg/ghg2.html (26 Nov. 2001)
Apps, M.J., Kurz, W.A., Beukema, S.J., and Bhatti, J.S. 1999. Carbon budget of the
Canadian forest product sector. Environmental Science and Policy No. 2. pg 25-41.
Elsevier Science Ltd.
Birdsey, R.A.. 1996. Chapter 1 – Carbon storage for major forest types and regions in the
conterminous United States. Forests and global change – Vol 2: Forest Management
Opportunities for Mitigating Carbon Emissions. Sampson, N., and Hair, D. (eds).
Washington, DC: American Forests.
European Environment Agency (EEA). 1999. EMEP/CORINAIR Atmospheric emission
inventory guidebook - Second ed. Copenhagen, Denmark.
http://reports.eea.eu.int/EMEPCORINAIR/en (26 Nov. 2001).
Fortum Power and Heat Oy. 2001. Methane and nitrous oxide emissions in the Finnish
energy production. Fortum, Filand: Fortum Power and Heat Oy, Technology.
International Energy Agency (IEA). 1998. The role of bioenergy in greenhouse gas
mitigation. Paris: International Energy Agency.
http://www.ieabioenergy.com/media/6_pospapa4.pdf (14 Feb. 2002).
Intergovernmental Panel on Climate Change (IPCC). 1996. Climate change 1995, the
science of climate change. New York: Cambridge University Press
Intergovernmental Panel on Climate Change (IPCC). 1997a. Revised 1996 IPCC guidelines
for national greenhouse gas inventories: Reporting instructions (Vol 1). IPCC National
Greenhouse Gas Inventory Program.
http://www.ipcc-nggip.iges.or.jp/public/gl/invs4.htm (26 Nov. 2001).
Intergovernmental Panel on Climate Change (IPCC). 1997b. Revised 1996 IPCC guidelines
for national greenhouse gas inventories: Workbook (Vol 2). IPCC National Greenhouse
Gas Inventory Program. http://www.ipcc-nggip.iges.or.jp/public/gl/invs5.htm (26 Nov.
2001).
Intergovernmental Panel on Climate Change (IPCC). 1997c. Revised 1996 IPCC guidelines
for national greenhouse gas inventories: Reference manual (Vol 3). IPCC National
Greenhouse Gas Inventory Program.
http://www.ipcc-nggip.iges.or.jp/public/gl/invs6.htm (20 Nov. 2001).
Intergovernmental Panel on Climate Change (IPCC). 2000a. Good practice guidance and
uncertainty management in national greenhouse gas inventories. IPCC National
Greenhouse Gas Inventory Program.
http://www.ipcc-nggip.iges.or.jp/public/gp/gpgaum.htm (26 Nov. 2001).

54
Intergovernmental Panel on Climate Change (IPCC). 2000b. IPCC special report on land
use, land-use change, and forestry. Cambridge University Press. pp 375.
http://www.grida.no/climate/ipcc/land_use/ (16 April 2002).
Intergovernmental Panel on Climate Change (IPCC). 2001. Third assessment report (TAR),
climate change 2001: The scientific basis. Working Group 1 Report, Technical
Summary. IPCC National Greenhouse Gas Inventory Program.
http://www.ipcc.ch/activity/tar.htm (31 October 2002).
Japan Paper Association (JPA). 2002. Various personal communications between JPA and
Reid Miner of NCASI. Tokyo: Japan Paper Association.
Loreti, C.P., Foster, S.A., and Obbagy, F.E. 2001. An overview of greenhouse gas emissions
verification issues. Cambridge, MA: A.D. Little.
http://www.pewclimate.org/projects/emissions_verification.cfm (28 Nov. 2001).
Loreti, C.P., Wescott, W.F., and Isenberg, M.A. 2000. An overview of greenhouse gas
inventory issues. Cambridge, MA: A.D. Little.
http://www.pewclimate.org/projects/greenhouse.cfm (28 Nov. 2001).
Matthews, R. 1996. The influence of carbon budget methodology on assessments of the
impacts of forest management on the carbon balance. Forest ecosystems, forest
management and the global carbon cycle. Apps and Price (eds.). Berlin Heidelberg:
Springer-Verlag.
Miner, R., and Upton, B. 2002. Methods for estimating greenhouse gas emissions from lime
kilns at kraft pulp mills. Energy 27(8):729-738.
National Council for Air and Stream Improvement, Inc. (NCASI). 1980. A study of wood-
residue fired power boiler total gaseous non-methane organic emissions in the Pacific
Northwest. Air Quality Technical Bulletin No.109. Research Triangle Park, NC:
National Council for Air and Stream Improvement, Inc.
National Council for Air and Stream Improvement, Inc. (NCASI). 1981. A study of the kraft
process lime kiln total gaseous non-methane organic emissions. Technical Bulletin
No. 358. Research Triangle Park, NC: National Council for Air and Stream
Improvement, Inc.
National Council for Air and Stream Improvement, Inc. (NCASI). 1985. Volatile organic
carbon emissions from wood residue fired power boilers in the Southeast. Technical
Bulletin No.455. Research Triangle Park, NC: National Council for Air and Stream
Improvement, Inc.
National Council for Air and Stream Improvement, Inc. (NCASI). 1999. Evaluation of the
EPA-recommended approach to predicting air emissions from pulp and paper landfills.
Technical Bulletin No.790. Research Triangle Park, NC: National Council for Air and
Stream Improvement, Inc.

55
Skog, K.E., and Nicholson, G.A. 1998. Carbon cycling through wood products: The role of
wood and paper products in carbon sequestration. Forest Products Journal 48(7/8).
Swedish Environmental Protection Agency. 2001. Sweden’s national inventory report 2001
– Submitted under the United National Convention on Climate Change. Stockholm:
Swedish Environmental Protection Agency.
http://www.environ.se/dokument/fororen/klimat/klimat/fcccdata/NIR2001.pdf (26 Nov.
2001).
United States Environmental Protection Agency (USEPA). 1995. Protocol for equipment
leak emission estimates, November 1995. Washington, DC: United States Environmental
Protection Agency. http://www.epa.gov/ttn/chief/efdocs/lks95_ch.pdf (24 Oct. 2002).
United States Environmental Protection Agency (USEPA). 1998. AP-42 emission factors
for municipal solid waste landfills – Supplement E, November 1998. Washington, DC:
United States Environmental Protection Agency.
http://www.epa.gov/ttn/chief/ap42/ch02/final/c02s04.pdf (26 Nov. 2001).
United States Environmental Protection Agency (USEPA). 2001. AP-42 emission factors
for wood residue combustion in boilers – Supplement G, July 2001. Washington, DC:
United States Environmental Protection Agency.
http://www.epa.gov/ttn/chief/ap42/ch01/final/c01s06.pdf (26 Nov. 2001).
Winjum, J.K., Brown, S., and Schlamadinger, B. 1998. Forest harvests and wood products:
sources and sinks of atmospheric carbon dioxide. Forest Science 44(2):274.
World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001a. The greenhouse gas protocol: A corporate accounting and reporting standard.
Washington, DC: World Resources Institute.
http://www.ghgprotocol.org/standard/ghg.pdf (26 Nov. 2001).
World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001b. Calculating CO2 emissions from the combustion of standard fuels and from
electricity/steam purchase - Guide to calculation worksheets. Washington, DC: World
Resources Institute. http://www.ghgprotocol.org/standard/stationarycombustion.doc
(26 Nov. 2001).
World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001c. Calculating CO2 emissions from the combustion of standard fuels and from
electricity/steam purchase - Calculation worksheets. Washington, DC: World Resources
Institute. http://www.ghgprotocol.org/standard/stationarycombustion.xls (26 Nov. 2001).
World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001d. Calculating CO2 emissions from mobile sources -Guide to calculation
worksheets. Washington, DC: World Resources Institute.
http://www.ghgprotocol.org/standard/mobile.doc (26 Nov. 2001).

56
ANNEX A

GREENHOUSE GAS EMISSIONS FROM STATIONARY FOSSIL FUEL


COMBUSTION – OVERVIEW OF METHODS IN EXISTING PROTOCOLS

1.0 IPCC – REVISED 1996 IPCC GUIDELINES FOR NATIONAL


GREENHOUSE GAS INVENTORIES AND MAY 2000 GOOD PRACTICES
DOCUMENT (IPCC 1997a, 1997b, 1997c, 2000)

1.1 Carbon Dioxide

The IPCC guidelines are designed for national-level reporting. Therefore, a certain amount
of judgment is involved in applying the guidelines at the company- or facility-level. In the
case of energy-related CO2 emissions, IPCC has established a three tiered system. Tier 1,
which is also part of the Reference Method for national reporting, uses fuel consumption
statistics and average CO2 emissions factors without regard to the types of combustion
technology employed. In the Reference Method, the Tier 1 approach is used to develop
national totals by applying it in a “top down” fashion, meaning that national fuel
consumption statistics are used. The Tier 1 approach can also used from the “bottom up” –
i.e., individual facilities can estimate CO2 emissions from fuel consumption data without
considering the type of combustion units being used (IPCC 1997c, pg. 1.7).

IPCC explains that “Tier 2 methods may be regarded as those dividing fuel consumption on
the basis of sample or engineering knowledge between technology types which are
sufficiently homogenous to permit the use of representative emission factors. Tier 3 methods
generally estimate emissions from activity figures (kilometre traveled or tonne-kilometer
carried, not fuel consumption) and specific fuel efficiency or fuel rates or, alternatively, using
an emission factor or factors expressed directly in terms of a unit of activity” (IPCC 1997c,
pg. 1.47).

Simply put, Tier 1 methods use emission factors that vary by fuel type only. Tier 2 methods
use emissions factors that vary by fuel and by type of combustion unit. Tier 3 methods use
emission factors that are based on a measure of activity other than fuel consumption (e.g., per
kilometer traveled).

In general, IPCC recommends using the most accurate and site-specific information
available. This suggests that, where possible, IPCC would recommend estimating energy–
related CO2 emissions by multiplying the quantities of fuel consumed by the actual carbon
content of the fuel and correcting for unoxidized carbon based on analysis of the ash and
emissions. While companies will normally have good records of fuel consumption, it is
unlikely that they will have data on the carbon content of all fuels they use. In addition, if
data exist on the amounts of unburned carbon in the ash and emissions, these data will
probably be sparse.

IPCC’s Tier 1 approach for estimating CO2 emissions from fossil fuel combustion uses
emission factors that vary by fuel. To reduce the uncertainty caused by using generic
information on the carbon contents of fossil fuels, IPCC prefers that the factors be expressed
as mass of carbon dioxide per unit of energy – e.g., tons carbon dioxide per terajoule of fossil

A-1
fuel burned. IPCC explains that “expressing the carbon emissions factor as the carbon
content per unit of energy released reduces [the] variation because of the close link between
the carbon content and energy value of the fuel.” In this report, unless shown otherwise,
fossil fuel energy content is always expressed as the net calorific value (NCV) of the fuel,
also known as the lower heating value (LHV).

Once the consumption statistics have been converted to units of energy, they must be
multiplied by the corresponding carbon emission factor for each fossil fuel. If a company has
site-specific data on the carbon content of the fossil fuels being burned, these data are
preferred. If such data are not available, IPCC accepts the use of generic carbon dioxide
emission factors. The default carbon dioxide emission factors for the IPCC reference method
are shown in Table 1-2 of the IPCC Guidelines Workbook. The emission factors for fuels
most likely to be of interest to pulp and paper mills are shown in the following table.

Table 1. IPCC Default Emission Factors for Fossil Fuels (after IPCC 1997b)
Converted from IPCC’s factors, which are expressed in units of carbon rather than carbon
dioxide
Fossil Fuel kg
CO2/TJ*
Crude Oil 73,300
Gasoline 69,300
Kerosene 71,900
Diesel Oil 74,100
Residual Fuel Oil 77,400
LPG 63,100
Petroleum Coke 100,800
Anthracite Coal 98,300
Bituminous Coal 94,600
Sub-Bituminous Coal 96,100
Lignite 101,200
Peat 106,000
Natural Gas 56,100
* assuming no unburned carbon

The IPCC Reference Manual also shows several equations that can be used to estimate the
carbon dioxide emission factors for natural gas and certain coals (IPCC 1997c).

For natural gas, IPCC also presents a relationship wherein the carbon emission factor is
shown as a function of the gross calorific value (HV) or high heating value of the natural gas
as follows.

A-2
Tons Carbon /TJ = 13.708 + (0.0828 x 10-3) x (HV in kJ/m3 – 37.234) (Eq.1)

Tons CO2 /TJ = Tons Carbon/TJ x 44/12

The 1996 IPCC Reference Manual also shows a relationship for estimating the carbon
emission factors for coals having a gross calorific value (HV) of between 31 and 37 terajoules
per kiloton on a dry mineral matter free (dmf) basis (IPCC 1997c).

Tons Carbon /TJ = 32.15 – (0.234 x HV in TJ/kiloton, dry mineral free basis) (Eq.2)

Tons CO2 /TJ = Tons Carbon/TJ x 44/12

After multiplying the amount of fuel burned (in units of energy) by the appropriate emission
factor (in units of ton carbon or tons CO2 per unit of energy), IPCC corrects the resulting
emission estimate to account for unoxidized carbon in the ash and emissions. IPCC
acknowledges that the amount of unoxidized carbon is highly variable between fuels and
combustion units. None-the-less, to correct CO2 emissions estimates for unburned carbon,
IPPC recommends default correction factors of 0.98 for coal, 0.99 for oil, and 0.995 for gas
(IPCC 1997c).

Where coal is burned one could estimate unoxidized carbon based on the carbon content of
the ash. IPCC acknowledges that for coal fired boilers most of the unoxidized carbon is in
the ash. This approach would not work, however, for units burning a combination of fossil
fuels and biomass fuels, such as those that burn coal (or other fossil fuels) in combination
with bark or wood waste. In these cases, it will probably not be possible to differentiate
between unoxidized carbon of fossil fuel origin and unoxidized carbon from the biomass.

IPCC discusses several studies that provide more resolution than the IPCC Tier 1 default
factors for unoxidized carbon. The IPCC Reference Manual lists several sources of data to
support Tier 2 approaches to estimating unoxidized carbon (although these data are included
in the Tier 1 section of the manual). An Australian study is cited which shows the
unoxidized carbon from stoker-fired boilers ranging from 1 to 12 percent for coals having ash
contents of 8 to 23 percent. A value of 2 percent is suggested as best practice, 5 percent for
average practice and 10 percent for worst practice (IPCC 1997c).

A similar study by British Coal produced the following data that can be used to develop Tier
2 estimates of unburned carbon as a function of combustion technology (IPCC 1997c).

Pulverized Coal – 1.6%


Traveling Grate Stoker – 2.7 to 5.4%
Underfeed Stoker – 4.0 to 6.6%
Shallow Bed Advanced Fluidized Bed Boiler – Up to 4.0%
Pressurized or Circulating Fluidized Bed Combustion – 3.0%

Finally, the IPCC Reference Manual indicates that well maintained natural gas fired boilers
often attain combustion efficiencies of 99.9 percent, suggesting that for most gas-fired
industrial boilers, the default value for unoxidized carbon (0.5%) may be too large (IPCC
1997c).

A-3
1.2 Methane

IPCC offers Tier 1 and Tier 2 approaches for estimating methane emissions from stationary
fossil fuel combustion.

For the “energy industries,” the Tier 1 emission factors are 1 kg methane/TJ for coal and
natural gas, and 3 kg methane/TJ for oil. For “manufacturing industries and construction,”
the emissions factors are 10 kg methane/TJ for coal, 5 kg/TJ for natural gas, and 2 kg/TJ for
oil (IPCC 1997c, pg. 1.35).

The Tier 2 factors are presented in two pieces – an uncontrolled emissions factor coupled
with a factor to reflect the effectiveness of various emission control devices. The
uncontrolled emissions factors are different for utility and industrial boilers. The
uncontrolled methane emission factors for industrial boilers are shown in Table 1-16 of the
IPCC Reference Manual. Table 1-17 lists methane emission factors for several industrial
processes involving fossil fuel combustion, including “lime kilns.” The factors for lime kilns
are intended for commercial lime kilns and may or may not be appropriate for kraft mill lime
kilns. The Tables are reproduced below. They also contain the emission factors for nitrous
oxide, discussed in more detail below. The industrial boiler and kiln emission controls
performance for methane is listed in the Reference Manual (Tables 1-21 and 1-22) as being
negligible, so the uncontrolled emissions in Tables 1-16 and 1-17 can be considered equal to
final controlled emissions (IPCC 1997c).

1.3 Nitrous Oxide

IPCC’s approach to estimating nitrous oxide emissions is the same as it uses for methane.
IPCC offers both Tier 1 and Tier 2 approaches. For both the “energy industries” and
“manufacturing industries and construction,” the Tier 1 emission factors are 1.4 kg nitrous
oxide/TJ for coal, 0.1 kg/TJ for natural gas, and 0.6 kg/TJ for oil. (IPCC 1997c, pg. 1.36)

The Tier two factors are presented in two pieces – an uncontrolled emission factor coupled
with a factor to reflect the effectiveness of various emission control devices. The
uncontrolled emissions factors are different for utility and industrial boilers. The
uncontrolled nitrous oxide emission factors for industrial boilers are shown in Table 1-16 of
the IPCC Reference Manual reproduced below. The Tier 2 approach involves correcting
uncontrolled emissions to account for the impact of control technologies. The only case
where IPCC provides the needed data, however, is in the case of selective catalytic reduction
applied to gas fired boilers. IPCC indicates that SCR accomplishes a 60% reduction in
uncontrolled nitrous oxide emissions from gas fired boilers. In all other cases, the
uncontrolled emissions must be used because there are no data supplied on control
technology effectiveness (shown as “not available” in Tables 1-21 and 1-22 of the Reference
Manual). (IPCC 1997c)

A-4
Table 2. IPCC Industrial Boiler Uncontrolled Emission Factors
(IPCC 1997c, Table 1-16)
Fuel and Equipment Type kg CH4/ kg N2O/
TJ TJ
Coal
Bit./Sub-bit. Overfeed Stoker Boilers 1 1.6
Bit./Sub-bit. Underfeed Stoker Boilers 14 1.6
Bit./Sub-bit. Hand-fed Units 87 1.6
Bituminous/Sub-bituminous Pulverized Dry Bottom, wall Fired 0.7 1.6
Bituminous/Sub-bituminous Dry Bottom, tangentially fired 0.7 0.5
Bituminous/Sub-bituminous Wet Bottom 0.9 1.6
Bituminous Spreader Stokers 1 1.6
Bit./Sub-bit. Fluidized Bed Combustor Circulating Bed 1 96
Bit./Sub-bit. Fluidized Bed Combustor Bubbling Bed 1 96
Anthracite Stokers NAV NAV
Anthr. Fluidized Bed Combustor Boilers NAV NAV
Anthracite Pulverized Coal Boilers NAV NAV
Oil
Residual Fuel Oil Boilers 3 0.3
Distillate Fuel Oil Boilers 0.2 0.4
LPG Boilers Propane NAV NAV
LPG Boilers Butane NAV NAV
Small Stationary Internal Comb. Engines Gasoline <250 hp (186 kW) NAV NAV
Small Stationary Internal Comb. Engines Diesel <600 hp (447 kW) NAV NAV
Large Stationary Diesel Engines >600 hp (447 kW) 0 NAV
Natural Gas
Large Boilers >100 MBtu/h (293 MW) 1.4 NAV
Small Boilers 10-100 MBtu/h (29.3-293 MW) 1.4 NAV
Heavy Duty Nat. Gas. Compressor Eng. – Turbines 0.6 NAV
Heavy Duty Nat. Gas. Compressor Eng. – 2-Cycle Lean Burn 17 NAV
Heavy Duty Nat. Gas. Compressor Eng. – 4-Cycle Lean Burn 13 NAV
Heavy Duty Nat. Gas. Compressor Eng. – 4-Cycle Rich Burn 2.9 NAV
Note: NAV means Not Available

Table 3. IPCC Uncontrolled Emission Factors for Commercial Lime Kilns


(IPCC 1997c, Table 1-17)
Fuel kg CH4/TJ fuel kg N2O/TJ fuel
Natural Gas 1.1 No Data
Oil 1.0 No Data
Coal 1.0 No Data

A-5
2.0 AUSTRALIA – NATIONAL GREENHOUSE GAS INVENTORY and
GREENHOUSE CHALLENGE WORKBOOK (Australian Greenhouse Office
1998, 1999, 2001a, 2001b)

There are a number of documents that provide helpful information on the approaches used in
Australia to develop estimates of GHG releases from pulp and paper mills. Some of these
documents include:

· Greenhouse Challenge Factors and Methods Workbook Version 3 (December 2001) (the
GCP Workbook) (AGO 2001b)
· National Greenhouse Gas Inventory (NGGI) Committee Workbook 1.1 for Fuel
Combustion Activities (Stationary Sources) (AGO 1998, 1999)
· APIC Environmental Performance Indicators (APIC 2002) (Contains information on
industry emissions)

Australia’s Greenhouse Challenge contains guidance for company-level reporting which is


more relevant for purposes of this report than national-level reporting. The following
material, therefore, is drawn largely from the tools issued by the Australian Greenhouse
Office for the Greenhouse Challenge program (AGO 2001b). Information on the methods
used to develop Australia’s national inventory are also available (AGO 2001a)

2.1 Carbon Dioxide

The Australian Greenhouse Office (AGO) Greenhouse Challenge Factors and Methods
Workbook contains emission factors for a variety of fossil fuels. Usually, two sets of factors
are shown – one for emissions from combustion, and another (full fuel cycle emission
factors) that also includes emissions from upstream operations (e.g., mining and fuel
processing). There is no mention in the workbook about correcting for unburned carbon, so
it is assumed that the emission factors have been adjusted to reflect this. In many cases, the
emission factors are for CO2 only but AGO indicates that “they can be used as CO2-
equivalent emission factors except for equipment with high CH4 or N2O emissions – notably
internal combustion engines, wood-fueled boilers and bagasse boilers.” A number of the
AGO Greenhouse Challenge workbook factors are shown in the Table 4 (AGO 2001b).

The Australian Paper Industry Council (formerly the Pulp and Paper Manufacturers
Federation of Australia) issues guidance for estimating greenhouse gas emission which is
based on Australia’s Greenhouse Challenge program (PPMFA 2001). The CO2 emission
factors used by APIC in 2001 are in Table 5 below.

A-6
Table 4. Australian Emission Factors for CO2 from Stationary Fuel Combustion
(Australian Greenhouse Office 2001b)
Fuel Direct combustion Full fuel cycle emissions q
emissions * kg CO2-equiv./GJ f
f
kg CO2/GJ
Black coal 90 NSW 96.0
Queensland 91.6
Brown coal 93 93.9
Brown Coal Briquettes 105 109.4
Coke 119.5 133.3
LPG: non transport 59.4 68.6
Kerosene 69.7 78.9
Heating Oil 69.7 78.9
Automotive diesel: non- 69.7 78.9
transport
Industrial/marine diesel fuel 69.7 78.9
Fuel oil 73.6 82.8
Natural Gas (large users) NSW & ACT 51.7 NSW & ACT 70.7
Victoria 51.6 Victoria 63.4
Queensland 53.0 Queensland 66.2
SA 51.6 SA 72.0
WA 53.1 WA 59.3
NT 51.7 NT 53.2
Australia Ave. 52.2 Australia Ave. 64.0
Waste Methane 51 56.0 (if used in internal
combustion engine)
Wood and wood waste 94.0 (d) 95.4 (if used in boiler)
Bagasse 96.8 (d) 98.2 (if used in boiler)
* The AGO indicates that these emission factors are for CO2 only but “they can be used
as CO2-equivalent emission factors except for equipment with high CH4 or N2O
emissions – notably internal combustion engines, wood-fuelled boilers and bagasse
boilers.”
f
Energy is expressed in Gross Calorific Value (High Heating Value)
q
Full Fuel Cycle emissions include upstream emissions from fuel production,
transportation etc.

A-7
Table 5. Emission Factors from the Pulp and Paper Manufacturers Federation of Australia
(PPMFA 2001) - Based on the Australia Greenhouse Challenge factors, but in different units
than shown in the table above
Full Fuel Cycle
Emission Factors q -
Unit of Energy Tons CO2-equiv. per
Fuel or Process or Fuel unit of energy or fuel
Electricity kWh Site Specific
Natural gas GJ Site Specific
LPG (non-transport) t 3.4026
Petroleum products
LPG transport kilolitres 1.8633
Aviation gasoline kilolitres 2.6977
Petrol kilolitres 2.8933
Automotive diesel oil kL 3.1806
Industrial/marine diesel kL 3.2630
fuel
Natural gas/CNG kL 2.3582
Coal t Site Specific
q
Full Fuel Cycle emissions include upstream emissions from fuel
production, transportation etc.

2.2 Methane and Nitrous Oxide

Although in many cases, the combustion emission factors shown in Table 4 are for CO2 only,
AGO indicates that “they can be used as CO2-equivalent emission factors except for
equipment with high CH4 or N2O emissions – notably internal combustion engines, wood-
fueled boilers and bagasse boilers.” These other sources are dealt with elsewhere. The Full
Fuel Cycle emission factors in the tables above include CO2, methane, and nitrous oxide
(AGO 2001b).

3.0 CHILE – FIRST NATIONAL COMMUNICATION TO THE CONFERENCE


OF THE PARTIES OF THE CONVENTION, OCTOBER 1999 (Chilean
National Environmental Commission 1999)

3.1 Carbon Dioxide

The 1999 National Communication indicates that “[t]he methodology used corresponds to
the Guidelines set forth by the IPCC and includes the changes introduced in 1996. The
emission coefficients used to calculate emissions were those default values proposed by the
IPCC for use when no coefficients were available that reflected the conditions in the
country.” The report goes on to say that “CO2 emissions were calculated two ways. The first
method uses the Reference Approach, which is based on determining the apparent
consumption of each fuel and calculating the corresponding CO2 emissions through

A-8
representative emission factors for each fuel, linked to the technology of the final user. The
second method is the final Use Approach, which is based on fuel consumption for each final
use and calculating CO2 emissions through the representative emission factors for each fuel ,
linked to the technology of the final user.” Accordingly, the methods used in Chile appear to
be a combination of IPCC Tier 1 and Tier 2 methods combined with default emission factors
(Chilean National Environmental Commission 1999).

3.2 Methane and Nitrous Oxide

The report indicates that for non-CO2 gases, the Final Use Approach and default emission
factors are used, as recommended by the IPCC. Thus, it appears that IPCC’s Tier 2 method
with default emission factors was used for methane and nitrous oxide. The report does not
list the emission factors used. In the summary tables, there are no methane or nitrous oxide
emissions shown for the pulp and paper sector. (Chilean National Environmental
Commission 1999)

4.0 FINLAND – ANNEX B OF “GREENHOUSE GAS EMISSIONS AND


REMOVALS IN FINLAND” (Technical Research Center of Finland 2001)

4.1 Carbon Dioxide

The Finnish report describes the methods as being consistent with the IPCC Tier 2 approach.
The emission factors as shown in the following table. Many of these factors are identical to
those used by IPCC but a number are from sources more appropriate to the situation in
Finland. The corrections for unoxidized carbon that are used to develop the Finnish
inventory are the same as the IPCC Tier 1 factors; i.e., 2% for coal, 1% for oil and oil
products, 1% for peat, and 0.5% for gas.

Table 6. CO2 Emission Factors for Fuel Combustion


(From Technical Research Center of Finland 2001, Table B-1)
Fuel ton CO2/TJ
Gasoline 73.5
Diesel 74.5
Light Fuel Oil 74.1
Residual Fuel Oil 77.4
Kerosene 73.5
Naphtha 72
LPG 63.1
Waste Oil 77.4
Hard Coal 94.6
Coke 108
Anthracite and Briquettes 94.6
Natural Gas 56.1
Peat 106

A-9
4.2 Methane

Finland’s report to UNFCCC indicates that “CH4 and N2O emission factors are based on
mainly national research (included evaluation and application of international research and
measurement data).... IPCC default values are used to a lesser degree.” The emission factors
given in the Annexes to the report, however, are only for small units (less than 1 MW)
(Technical Research Center of Finland 2001).

4.3 Nitrous Oxide

Finland’s report to UNFCCC indicates that “CH4 and N2O emission factors are based on
mainly national research (included evaluation and application of international research and
measurement data).... IPCC default values are used to a lesser degree.” The emission factors
given in the Annexes to the report, however, are only for small units (less than 1 MW)
(Technical Research Center of Finland 2001).

The Finland report also indicates that “indirect emissions due to atmospheric deposition of
NH3 and NOX emissions…are calculated…for energy-related and industrial NOX emissions
using the IPCC methodology and default emission factor (0.01 kg N2O-N/kg NOX-N
emitted). The indirect N2O emissions caused by nitrogen deposition due to NOX emissions
in the energy sector are included in the emission estimates for the relevant sectors”
(Technical Research Center of Finland 2001). This does not appear to be required by the
IPCC estimation methods for energy production, although it is mentioned in the IPCC’s
methods for estimating GHG releases from agriculture. (See Section 4.5.4.A and Table 4-23
in IPCC 1997c.)

5.0 CANADA – “GREENHOUSE GAS ACTION PLAN GUIDELINES – 2001”


(FPAC 2001) AND “CANADA’S GREENHOUSE GAS INVENTORY, 1997
EMISSIONS AND REMOVALS WITH TRENDS” (Environment Canada 1999)

5.1 Carbon Dioxide

The Forest Products Association of Canada (FPAC) Guidelines are intended to facilitate
forest company participation in the Voluntary Callenge and Registry Program (VCR) for
GHGs and the Canadian Industry Program for Energy Conservation/Industrial Energy
Innovators (CIPEC/IEI) for energy conservation, as well as several Provincial programs. The
method for estimating CO2 emissions from stationary fossil fuel combustion is generally
comparable to IPCC’s Tier 1 sectoral approach although the FPAC and VCR guidelines offer
no emission factors for coal. Instead the user is instructed to estimate emissions based on the
carbon content of the coal being burned. The emission factors, which are published in
Appendix A of the FPAC Guidelines, are shown below (FPCA 2001). Like the factors used
to develop Canada’s inventory, the emission factors published by FPAC are normalized to
the volume or weight of fuel rather than heat content.

The Canadian Inventory document contains CO2 emission factors for coal that vary by type
of coal and region of the country. There is no correction factor applied for unoxidized
carbon, this having apparantly been incorporated into the emission factors.

A - 10
5.2 Methane

Methods for estimating methane emissions from fossil fuel combustion are comparable to
IPCC’s Tier 1 methods except that (a) the emission factors are mass- or volume-based rather
then energy-based, and (b) there is no discussion of a correction for the effectiveness of
control devices (an unimportant difference since IPCC assumes zero percent removal of
methane).

5.3 Nitrous Oxide

Like methane, methods for estimating N2O emissions from fossil fuel combustion are
comparable to IPCC’s Tier 1 methods except that (a) the emission factors are mass- or
volume-based rather then energy-based, and (b) there is no discussion of a correction for the
effectiveness of control devices (an unimportant difference since IPCC provides removal
data only for gas fired boilers equipped with selective catalytic reduction).

Table 7. Emission Factors from Appendix A of FPAC 2001


Fuel CO2 CH4 N2O
Natural Gas 1,880 g/m3 0.048 g/m3 0.02 g/m3
Light (distillate) oil 2,830 g/l 0.006 g/l 0.013 g/l
Heavy (residual) oil 3,090 g/l 0.12 g/l 0.013 g/l
Coal (conventional Use carbon content of 0.015 g/kg 0.11 g/kg
industrial boiler) coal to determine
Coal (fluidized bed) Use carbon content of 0.015 g/kg 2.11 g/kg
coal to determine
Propane Heaters 1,530 g/l 0.03 g/l -
Diesel Motors (stationary) 2,730 g/l 0.26 g/l 0.40 g/l

6.0 FRANCE – “INVENTAIRE DES ÉMISSIONS DE GAZ À EFFET DE SERRE


EN FRANCE DE 1990 À 1999” [INVENTORY OF GREENHOUSE GAS
EMISSIONS IN FRANCE FOR THE PERIODE 1990-1999] (CITEPA 2000)

6.1 Carbon Dioxide

The French government starts with CORINAIR methods (EEA 1999) to develop a more
detailed inventory than required by UNFCCC and then consolidates the results into sector-
level and national reports in the UNFCCC Common Reporting Framework. The IPCC 1996
Revised Guidelines are largely harmonized with CORINAIR’s.

Carbon dioxide emissions are estimated using source measurements if available. If not, an
approach similar to IPCC’s Tier 2 approach is used. National emissions are also estimated
using IPCC’s Tier 1 approach and the results compared to those resulting from the combined
Tier 2 results. The emission factors for fossil fuels are shown the table below.

A - 11
6.2 Methane and Nitrous Oxide

Site specific measurements are used where available. In the absence of measured data,
emission factors are used, many of which reflect the situation in France. The emission
factors for CH4 and N2O are not included in the report.

Table 8. French Emission Factors for CO2 from Fossil Fuels


from CITEPA 2000
Fuel tons CO2 / TJ
coking coal or thermal coal 95
sub-bituminous coal 96
lignite 100
coke 107
coke from lignite 108
petroleum coke 96
heavy fuel 78 + / - 1 according to quality
gas oil 75
kerosene and jet fuel 74
gasoline 73
natural gas 57
LPG 64
refinery gas 56
coke oven gas 47
blast furnace gas 268

7.0 JAPAN – INFORMATION FROM THE JAPAN PAPER ASSOCIATION (JPA


2001)

7.1 Carbon dioxide, methane, and nitrous oxide

In Japan, the members of the industry group Keidanren or the Federation of Economic
Organization have committed to voluntary reductions in GHG emissions. The information
for the Japanese pulp and paper industry is collected and submitted to Keidanren by the
Japan Paper Association. Only CO2 is included in the reporting, not CH4 and N2O.

The Japan Ministry of Environment also has a program wherein the GHG emissions for the
country are estimated annually. This program includes CO2 (not including biomass) and
N2O. The emission factors shown below are from the Ministry of Environment program.

A - 12
Table 9. Japan Ministry of the Environment Emission Factors for Fossil Fuels
Combustion Fuel CO2 N2O
Unit
Boiler Crude Oil 2.65 kg/litre 0.00000052 kg/litre
Boiler Kerosene 2.51 kg/litre
Boiler Light Oil 2.64 kg/litre -
Boiler Heavy Fuel Oil 2.96 kg/litre 0.00000056 kg/litre
Boiler Coal 2.37 kg/kg 0.00012 kg/kg
Boiler Petroleum Coke 3.3 kg/kg 0.00013 kg/kg
Boiler Town Gas 2.15 kg/kg -
Turbine Crude Oil 2.65 kg/litre 0.0000011 kg/litre
Turbine Kerosene 2.51 kg/litre 0.000001 kg/litre
Turbine Light Oil 2.64 kg/litre 0.0000011 kg/litre
Turbine LPG 3.02 kg/kg 0.0000014 kg/kg
Turbine LNG 2.79 kg/kg 0.0000015 kg/kg
Turbine Town Gas 2.15 kg/kg 0.0000013 kg/m3
Diesel engine Crude Oil 2.65 kg/litre 0.000062 kg/litre
Diesel engine Kerosene 2.51 kg/litre 0.000062 kg/litre
Diesel engine Light Oil 2.64 kg/litre 0.000062 kg/litre

8.0 SWEDEN – SWEDEN’S NATIONAL INVENTORY REPORT (Swedish


Environmental Protection Agency 2001)

8.1 Carbon Dioxide

In its annual report to the UNFCCC, the methods used by Sweden in reporting CO2
emissions from steam and power generation are essentially Tier 1 IPCC 1996 methods. The
CO2 emission factors used by Sweden are found in Table 1.7.2 of “Sweden’s National
Inventory Report 2001” and the emission factors for the fuels of potential interest to pulp and
paper are shown below.

A - 13
Table 10. Swedish Emission Factors for CO2 from Fossil Fuels
From Table 1.7.2 in Swedish Environmental Protection Agency 2001
Fuel Emission factor, g CO2/MJ
Gas oil and diesel oil 75.3
Residual fuel oil 76.2
Liquefied petroleum gas 65.1
Natural gas 56.5
Coke oven coke 103.0
Peat (industry consumption) 97.1
Gasoline 72.6
Other kerosene 73.1

8.2 Methane

Sweden uses an IPCC Tier 1 approach for estimating methane emissions. The emission
factors are contained in Table 1.7.3 of its 2001 Inventory Report. The table is reproduced
below.

Table 11. Swedish Emission Factors for CH4 from Fossil Fuels
From Table 1.7.3 in Swedish Environmental Protection Agency 2001
Fuel Emission factor, g CH4/MJ
Gas oil and diesel oil 0.001
Residual fuel oil 0.002
Liquefied petroleum gas 0.001
Natural gas 0.001
Coke oven coke 0.002
Peat (industry consumption) 0.03

8.3 Nitrous Oxide

Like CO2 and methane, Sweden uses an IPCC Tier 1 approach for estimating nitrous oxide
emissions from fossil fuel fired units producing steam and/or power. The emission factors
are in Table 1.7.4 of its 2001 inventory. The table is reproduced below. The factors are not
identified as being uncontrolled, so it is assumed that they reflect actual (i.e., controlled)
emissions.

A - 14
Table 12. Swedish Emission Factors for N2O from Fossil Fuels
From Table 1.7.4 in Swedish Environmental Protection Agency 2001
Fuel Emission factor, g N2O/MJ
Gas oil and diesel oil 0.002
Residual fuel oil 0.005
Liquefied petroleum gas 0.002
Natural gas 0.002
Coke oven coke 0.02
Peat (industry consumption) 0.01

9.0 UNITED KINGDOM – UK GREENHOUSE GAS INVENTORY, 1990 – 1999


(AEA Technology 2001)

9.1 Carbon Dioxide

The UK approach differs from the IPCC Tier 1 approach in several respects. First, the
emission factors for solid and liquid fuels are based on mass of fuel rather than energy
content. Second, the emission factors for gaseous fuels are based on gross calorific value
(GCV) rather than net calorific value (NCV). Third, the emission factors incorporate the
impact of unoxidized carbon, so no correction is needed. Finally, the UK approach has
different emission factors for a single fuel used in different sectors of the economy (AEA
Technology 2001).

The emission factors used by the UK in developing its inventory are included in Tables 1 to 4
of Appendix 2 to “UK Greenhouse Gas Inventory, 1990-1999.” These tables are reproduced
below (AEA Technology 2001).

9.2 Methane

The differences between IPCC’s Tier 1 approach and the UK approach for estimating CO2
emissions are described above. These differences also hold true for estimation of methane
from fossil fuel combustion in steam and power generation. It appears that the emission
factors represent final emissions (i.e., following control technologies). The emission factors
used by the UK in developing its inventory are included in Tables 1 to 4 of Appendix 2 to
“UK Greenhouse Gas Inventory, 1990-1999.” The emission factors from these tables are
shown below (AEA Technology 2001).

9.3 Nitrous Oxide

The differences between IPCC’s Tier 1 approach and the UK approach for estimating CO2
emissions are described above. These differences also hold true for estimation of nitrous
oxide from fossil fuel combustion in steam and power generation. It appears that the UK
emission factors represent final emissions (i.e., following control technologies). From a
practical standpoint, this differs from the IPCC approach only to the extent that IPCC
provides a separate factor to correct uncontrolled emissions of nitrous oxide from gas boilers

A - 15
equipped with selective catalytic reduction. The emission factors used by the UK in
developing its inventory are included in Tables 1 to 4 of Appendix 2 to “UK Greenhouse Gas
Inventory, 1990-1999.” The emission factors from the tables are shown below (AEA
Technology 2001).

Table 13. UK Emission Factors


From Tables 1 through 4 of Appendix 2 of AEA Technology 2001
Fuel Source CH4 Emissions N2O Emissions
Burning Oil Misc. Sources 0.0432 kg/t 0.0265 kg/t
Gas Oil Agriculture 0.225 kg/t 0.026 kg/t
Gas Oil Industry (General) 0.0432 kg/t 0.026 kg/t
Fuel Oil Industry (General) 0.122 kg/t 0.0243 kg/t
Coal Industry (General) 0.011 kg/t 0.213 kg/t
Coal Autogenerators 0.02 kg/t 0.0658 kg/t
Coke Industry (General) 0.011 kg/t 0.221 kg/t
Solid Smokeless Fuel (SSF) Industry (General) 0.011 kg/t 0.218 kg/t
LPG Industry (General) 0.0971 g/therm* 0.0095 g/therm*
Natural Gas Industry (General) 0.38 g/therm* 0.0095 g/therm*
Natural Gas Autogenerators 0.579 g/therm* 0.351 g/therm*
* Based on Gross Calorific Value

10.0 UNITED STATES – INVENTORY OF U.S. GREENHOUSE GAS EMISSIONS


AND SINKS: 1990-1999 (USEPA 2001a)

10.1 Carbon Dioxide

The methods used by the US EPA to develop its annual inventory for CO2 from stationary
fossil fuel combustion are similar to the IPCC Tier 1 methods. One difference is that the US
fossil fuel consumption statistics are shown in GCV rather than NCV. (The EPA report
indicates that NCV is used for certain biomass fuels.)

Another departure from the default IPCC Tier 1 approach regards the unoxidized carbon in
coal burning – US EPA uses 1% rather than the 2% default value suggested by IPCC.

The carbon factors and correction factors for unoxidized carbon are contained in Table A-13
in Annex A of the EPA report. The emission factors from the table are shown below.
Although the report is not explicit, it can be assumed that the carbon factors are based on
gross calorific values (USEPA 2001a).

A - 16
Table 14. US EPA CO2 Emission Factors for Fossil Fuels
From Table A-13 from Annex A of USEPA 2001a
(Heat content in Gross Calorific Value)
Fuel Type Tg C/QBtu HHV Tg CO2/QBtu HHV
Industrial Other Coal* 25.63 93.98
Coke Imports 27.85 102.12
Utility Coal* 25.76 94.45
Natural Gas 14.47 53.06
Distillate Fuel Oil 19.95 73.15
Kerosene 19.72 72.31
LPG* 16.99 62.30
LPG (energy use)* 17.11 62.74
LPG (non-energy use)* 16.88 61.89
Motor Gasoline* 19.36 70.99
Residual Fuel 21.49 78.80
* EPA indicates these vary from year to year. The values shown are for
1999.

10.2 Methane

EPA used IPCC’s Tier 1 default methodology and emission factors to estimate methane
emissions from fossil fuel combustion in steam and power production (USEPA 2001a).

10.3 Nitrous Oxide

EPA used IPCCs Tier 1 default methodology and emission factors to estimate nitrous oxide
emissions from fossil fuel combustion in steam and power production (USEPA 2001a).

11.0 WRI/WBCSD – THE GREENHOUSE GAS PROTOCOL AND SUPPORTING


GUIDANCE (WRI 2001a, 2001b, 2001c)

11.1 Carbon Dioxide

WRI/WBCSD recommends that site-specific information on fuel use and fuel carbon content
(if available) be used to estimate CO2 emissions from stationary combustion sources. If site
specific data on fuel carbon content are unavailable, the WRI/WBCSD guidance material
contains a large number of emission factors from other sources. Some of the ones of
potential interest to pulp and paper mills are shown in the following tables. The guidance
indicates that WRI/WBCSD protocol does not require a correction for unoxidized carbon but
notes that this may not be accurate for small coal fired units (WRI 2001b, 2001c).

A - 17
11.2 Methane and Nitrous Oxide

The WRI/WBCSD guidance does not include CH4 and N2O. In the words of the October
2001 guidance, “a much greater effort is required to estimate CH4 and N2O emissions from
stationary sources [compared to CO2], and a much higher level of uncertainty exists.
Consequently, due the relative insignificance of CH4 and N2O emissions from most
stationary sources, [the WRI/WBCSD protocol] only includes guidance for estimating CO2
emissions from stationary sources” (WRI 2001b).

A - 18
Table 15. WRI/WBCSD CO2 Emission Factors for Stationary Combustion, October 2001 (WRI 2001c)
Fuel type kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 / kg CO2 /
GJ fuel million btu MWh fuel GJ fuel million btu MWh fuel metric short tons litres fuel gallons fuel standard 1000 SCF
LHV fuel LHV LHV HHV fuel HHV HHV tons fuel fuel cubic meters

Distillate fuel (No.1, No.2, No.4 fuel 74.01 78.08 266.41 69.38 73.20 249.83 3142 (UK 2850 (UK 2.68 10.15
oil and diesel) DETR) DETR)
Residual fuel oil (No.5, No.6 fuel oil) 77.30 81.55 278.26 74.77 78.88 269.22 3117 (UK 2828 (UK 3.12 11.81
DETR) DETR)
LPG 63.20 66.68 227.50 59.78 63.07 215.26 1.54 5.81
Propane 62.99 66.45 (EIA) 226.8 (EIA) 59.84 63.13 215.46 1.52 5.75
(EIA)
Natural gas (dry) 56.06 59.14 201.80 50.34 53.11 181.26 1.93 54.70
Anthracite 98.30 103.70 353.85 97.77 103.15 352.05 1926.04 1747.30
Bituminous coal 94.53 99.73 340.28 88.27 93.12 317.82 2465.61 2236.80
Sub-bituminous coal 96.00 101.28 345.57 91.45 96.48 329.28 1857.91 1685.50
Lignite 101.12 106.68 364.00 92.61 97.70 333.45 1395.83 1266.30
Peat 105.89 111.71 381.26 100.6 106.12 362.2
(IPCC) (IPCC) (IPCC)
Petroleum coke 100.76 106.30 362.71 96.80 102.12 348.53 3384.37 3070.30 3.88 14.69
Coke oven / gas coke 108.09 114.03 389.18 102.68 108.33 369.72
(IPCC) (IPCC) (IPCC)
Sludges 79.90 84.29 287.68 75.90 80.08 273.30
Pitch 79.90 84.29 287.68 75.90 80.08 273.30
Saw dust impregnated 75.10 79.23 256.31 71.34 75.27 243.49
Tires and tire derived fuel 85.78 90.49 308.86 81.49 85.97 293.41 3080.03 2794.20
Wood and wood waste 100.44 105.97 361.67 95.42 100.67 343.58 1906.97 1730.00
(EIA) (EIA) (EIA)
Sources: Emission factors based on higher heating values from EIA, For fuels not covered by IPCC, emission factors were
2001, Appendix B, if not otherwise indicated calculated based on EIA values using the following formula:
Emission factors based on lower heating values from IPCC, Emission factor (LHV) = Emission factor (HHV) / 0.95 for
1999, Volume 2, Section 1, if not otherwise indicated solid/liquid fuels and
Emission factors based on volumetric or mass units from EIA, Emission factor (LHV) = Emission factor (HHV) / 0.90 for
2001, Appendix B, if not otherwise indicated gaseous fuels
Other sources used: For fuels not covered by EIA, emission factors were
DETR, 1999, Section 4.2 (DETR factors are specifically calculated based on IPCC values using the same conversion
labeled in the table) factor between lower and higher heating values
Holderbank, 2000 (Holderbank factors are in italics)

A - 19
ANNEX B

GREENHOUSE GAS EMISSIONS FROM


KRAFT MILL LIME KILNS AND CALCINERS

There has been considerable confusion about the correct way to estimate GHG emissions
from kraft mill lime kilns and calciners. Some of the approaches suggested by the protocols
included in this review are presented in this section. In addition, this section contains a
discussion of the correct approach for estimating these emissions. The correct approach is to
include only CO2 emissions associated with the fossil fuels burned in the kiln or calciner, and
CH4 and N2O emissions if deemed material to the inventory.

1.0 EXISTING PROTOCOLS

Environment Canada, in “Canada’s Greenhouse Gas Inventory, 1997 Emissions and


Removals with Trends,” (page 41) states that “[e]missions from the regeneration of lime
from spent pulping liquors at pulp mills are not included in the inventory. Since this CO2 is
biogenic in origin, it is recorded as a change in forest stock in the ‘Land Use Change and
Forestry’ category” (Environment Canada 1999).

FPAC reiterates this position in its “2001 Greenhouse Gas Action Plan Guidelines” (page 11)
where it states that “…emissions from the regeneration of lime from spent pulping liquors at
pulp mills should not be reported, as this CO2 is biogenic in origin” (FPAC 2001).

IPCC appears to imply that CO2 emissions from kraft mill lime kilns and calciners should be
included GHG inventories when it states that “[i]ndustries that regenerate lime from waste
calcium carbonates (e.g., wood pulp and paper plants) are unlikely to report their lime
production. Omission of these data may lead to an underestimation of lime production for a
country…” (IPCC 2000, page 3.23). The question of whether CO2 should be included,
however, is not directly addressed. IPCC suggests emission factors of 1.0 and 1.1 kg CH4/TJ
for oil-fired and gas-fired lime kilns, respectively (IPCC 1997c, Table 1-17). These factors
were not developed for kraft mill lime kilns, however.

EPA appears to be uncertain about the situation. In “U.S. Greenhouse Gas Emissions and
Sinks: 1990-1999” (page 3-8) EPA states that “[i]n the paper industry, mills that employ the
sulfate process (i.e., Kraft) consume lime in order to causticize a waste sodium carbonate
solution (i.e., black liquor). Most sulfate mills recover the waste calcium carbonate after the
causticizing operation and calcine it back into lime -thereby generating CO2 -for reuse in the
pulping process. However, some of these mills capture the CO2 released in this process to be
used as precipitated calcium carbonate (PCC). Further research is necessary to determine to
what extent CO2 is released to the atmosphere through generation of lime by paper mills”
(USEPA 2001a).

B-1
2.0 CORRECTLY ACCOUNTING FOR GREENHOUSE GAS EMISSIONS
FROM KRAFT MILL LIME KILNS AND CALCINERS

2.1 Carbon Dioxide

NCASI recently examined the question of GHG emissions from lime kilns (Miner and Upton
2002). Much of the following discussion is taken from that paper.

The recovery and reuse of pulping chemicals is essential to the kraft pulping process. The
recovery process involves two interconnected loops, depicted in Figure 1. These can be
thought of as a sodium loop and a calcium loop. In the sodium loop, white liquor, a mixture
of sodium hydroxide and sodium sulfide, is added to wood chips in a pulping digester. The
pulping process results in most of the non-fibrous material being dissolved from the wood.
The digester is discharged at the end of the pulping process, yielding a combination of wood
fiber and a spent pulping liquor called black liquor because of the very dark color imparted
by the dissolved lignin-derived material. The black liquor, which also contains the spent
pulping chemicals, is separated from the wood fibers by washing, concentrated by
evaporation, and then sent to a recovery furnace where it is burned under controlled
conditions.

The recovery furnace produces large amounts of steam which are used throughout the mill.
In addition, due to the controlled oxygen deficient conditions in the bottom of the furnace, a

B-2
molten smelt consisting primarily of sodium carbonate and sodium sulfide is produced. The
sodium sulfide is formed in the recovery furnace by the reduction of a variety of sulfur
compounds in the black liquor. The sodium carbonate is formed by the reaction of sodium
compounds (primarily sodium oxide and sodium sulfide) with CO2, a product of combustion
of the wood-derived material in the black liquor (i.e., biomass).

Molten smelt is discharged from the bottom of the recovery furnace and dissolved in water to
produce green liquor. Subsequently, the sodium carbonate in the green liquor is converted
into sodium hydroxide by reacting with calcium hydroxide in causticizers, producing a
calcium carbonate precipitate. White liquor, the combination of sodium hydroxide and
sodium sulfide needed for pulping, is produced by removing the precipitate. In this way the
sodium loop is closed. Because small amounts of sodium are lost, mills commonly make up
this loss by adding sodium carbonate, sodium sulfate, or sodium hydroxide, depending on
whether the mill needs the additional sulfur and the capacity of various processes in the
recovery system.

The calcium loop intersects the sodium loop at the causticizers. In the causticizers, calcium
carbonate (CaCO3) is formed by the reaction of calcium hydroxide with the sodium
carbonate in green liquor. The carbon contained in this calcium carbonate originated in the
wood, was converted to biomass CO2 in the recovery furnace, subsequently reacted with
sodium salts in the recovery furnace to form sodium carbonate, and was finally converted to
calcium carbonate in the causticizers. In essence, therefore, the reactions in the causticizers
accomplish a transfer of biomass carbon from the sodium loop to the calcium loop.

The calcium carbonate formed in the causticizers is separated from the white liquor,
dewatered (forming a material called lime mud), and washed before being burned in a lime
kiln or calciner to produce calcium oxide. The CO2 liberated in the conversion of calcium
carbonate to calcium oxide in the lime kiln contains carbon which originated in wood and
was transported to the calcium loop via the sodium carbonate in green liquor.

Therefore, except in cases where CaCO3 is added to the recovery system as a make-up
chemical, the only CO2 that should be included in GHG inventories is that amount from
fossil fuels burned in the kiln or calciner. Elsewhere in this report, there is additional
discussion of GHG emissions that may sometimes be related to carbonate-containing make-
up chemicals used by some mills.

2.2 Methane

Small amounts of methane have been found in lime kiln emissions at some kraft mills.
Sampling at three mills in the late 1970s found that methane concentrations were usually less
than 1 ppm by volume, although values has high as 34 ppm were measured (NCASI 1981).
Of 73 measurements made at the three mills, all but 7 were less than 5 ppm. For these three
mills, 5 ppm corresponded to approximately 0.004 kg methane per metric ton of pulp. At an
assumed lime kiln fuel consumption rate of 1.5 GJ/ton pulp, this is equal to 2.7 kg CH4/TJ.

B-3
IPCC suggests emission factors of 1.0 and 1.1 kg CH4/TJ for oil-fired and gas-fired lime
kilns, respectively. (IPCC 1997c, Table 1-17). These factors were not developed for kraft
mill lime kilns, however.

Although the data are limited and old, the NCASI emission factor of 2.7 kg CH4/TJ is
probably more appropriate for estimating methane emissions from kraft mill lime kilns than
the IPCC emission factors for commercial kilns. The emission factor is likely to be revised,
however, as new data are generated.

2.3 Nitrous Oxide

Because nitrous oxide is formed in some combustion processes, it is appropriate to examine


the potential for nitrous oxide emissions from kraft mill lime kilns. This gas is potentially
important because it is usually assumed to have a greenhouse gas global warming potential
310 times greater than CO2. IPCC has reviewed the literature and has concluded that
formation of nitrous oxide is unlikely outside a range of combustion temperatures of
approximately 538°C to 927°C (1000oF to 1700°F) (IPCC 1997c). The calcination of lime
commences at approximately 816°C (1500oF) and, in a lime kiln, normally involves
temperatures in the range of 980°C to 1200°C (1800oF to 2200oF) (Hough 1985). Heating
calcium solids to these temperatures in a lime kiln requires combustion temperatures above
the range thought to be suitable for nitrous oxide formation. Kraft mill lime kilns, therefore,
are not expected to be a significant source of nitrous oxide.

Some mills use calciners instead of kilns to regenerate lime. Because calciners operate at
lower temperatures (maximum temperature of about 870°C or 1600°F) there appears to be a
potential for N2O generation, but data are lacking (Hough 1985).

It is reasonable, therefore, to assume that N2O emissions from lime kilns are so low that they
need not be reported. In the case of calciners, however, this assumption may not be
reasonable, but there are no data so it is suggested that the N2O emission factors for
comparable-sized fossil fuel boilers be used (see Annex A).

2.4 Use of Kraft Mill Lime Kiln Gases to Manufacture Precipitated Calcium
Carbonate

Calcium carbonate pigment can be used as a coating and filler material in the production of
some grades of paper and paperboard. Calcium carbonate pigment is manufactured by
grinding limestone or marble, or by chemical precipitation. At some mills, the correct
characterization of lime kiln emissions is complicated by the now-common practice of using
the lime kiln stack gas CO2 (or gas from another boiler) to manufacture precipitated calcium
carbonate (PCC) at satellite plants. The PCC manufacturing process involves the reaction of
CO2-rich lime kiln gas with purchased calcium oxide (lime) to produce PCC. Although other
mill stack gases are sometimes used, the lime kiln stack is favored, primarily because of its
higher CO2 content.

While the amounts of CO2 used in PCC manufacturing at satellite plants appear to be small
compared to the industry’s overall emissions, the quantities can be significant at individual

B-4
mills because 50% or more of the CO2 in the stack gas can be consumed in the PCC
manufacturing process. Indeed, some kraft mills use biomass to supply almost all of their
energy needs, with the only significant use of fossil fuel being the lime kiln. At such mills,
the amounts of CO2 captured in PCC manufacture can actually exceed the amounts of fossil
CO2 emitted by the mill.

Because the CO2 exported to PCC plants is not emitted by the mill, nor by any source owned
by the company, this CO2 is reported as an export rather than an emission. Where CO2 is
exported from lime kilns, the gas consists of a combination of fossil fuel-derived CO2 and
climate-neutral biomass-derived CO2. In the suggested reporting format for the calculation
tools, these two types of CO2 exports are reported separately. The data on exports of fossil
fuel-derived CO2 are reported with other data on fossil fuel-derived CO2 and the exports of
biomass-derived CO2 are covered in the “Supporting Information on Biomass” supplied in
Annex I.

With respect to PCC manufactured from mill GHG emissions, it is important to note that the
calculation tools are intended to help characterize a facility’s or company’s greenhouse gas
emissions and not the fate of these emissions. Nor are they intended to address the life cycle
tradeoffs associated with use of mill emissions as a raw material in PCC manufacture. These
are questions that require a much broader analysis than is possible within the scope of this
inventory guidance.

B-5
ANNEX C

CH4 AND N2O EMISSIONS FROM BIOMASS COMBUSTION – OVERVIEW OF


METHODS IN EXISTING PROTOCOLS

1.0 IPCC – REVISED 1996 IPCC GUIDELINES FOR NATIONAL


GREENHOUSE GAS INVENTORIES AND MAY 2000 GOOD PRACTICES
DOCUMENT (IPCC 1997a, 1997b, 1997c, 2000)

1.1 Methane

IPCC offers Tier 1 and Tier 2 approaches for estimating methane emissions from biomass
combustion. For the “energy industries” and the “manufacturing industries and
construction,” the Tier 1 methane emission factor is 30 kg CH4/TJ for wood, wood waste,
and other biomass and wastes (IPCC 1997c).

The Tier two factors are presented in two pieces – an uncontrolled emissions factor coupled
with a factor to reflect the effectiveness of various emission control devices. The only
methane emission factors for biomass, however, are for industrial boilers. Table 1-16 of the
IPCC Reference Manual describes emission factors four types of wood burning, fuel
cell/Dutch oven boilers, stoker boilers, FBC boilers, and bagasse/Agricultural waste boilers.
The only one of these with a corresponding methane emission factor is stoker boilers where
the emission factor is 15 kg CH4/TJ. The methane removal accomplished by industrial
boiler emission controls is listed in the IPCC Reference Manual (Table 1-21) as being
negligible, so the uncontrolled emissions are the same as final controlled emissions (IPCC
1997c).

For both Tiers 1 and 2, IPCC appears to have relied, at least in part, on the 1995 EPA AP-42
document, but the methane and nitrous oxide factors for biomass have been changed in more
recent versions of AP-42. See the discussion of the US protocol below for the most recent
AP-42 emission factors.

1.2 Nitrous Oxide

IPCC’s approach to estimating nitrous oxide emissions is the same as it uses for methane.
IPCC offers both Tier 1 and Tier 2 approaches, but includes emission factors only for Tier 1.
For the “energy industries” and the “manufacturing industries and construction,” the Tier 1
nitrous oxide emission factor is 4 kg N2O/TJ for wood, wood waste, and other biomass and
wastes (IPCC 1997c).

Table 1-16 of the IPCC Reference Manual, which contains Tier 2 emission factors, discusses
emission factors for four types of wood burning, fuel cell/Dutch oven boilers, stoker boilers,
FBC boilers, and bagasse/agricultural waste boilers, but there are no factors in the table for
N2O (IPCC 1997c).

C-1
For both Tiers 1 and 2, IPCC appears to have relied, at least in part, on the 1995 EPA AP-42
document, but the methane and nitrous oxide factors for biomass have been changed in more
recent versions of AP-42. See the discussion of the US protocol below for the most recent
AP-42 emission factors.

2.0 AUSTRALIA (AGO 1998, 1999, 2001a, 2001b)

2.1 Methane and Nitrous Oxide

The Factors and Methods Workbook issued by the Australian Greenhouse Office does not
include separate emission factors for methane and nitrous oxide (AGO 2001b). Emission
factors for methane and nitrous from industrial wood combustion are published, however, in
the Energy Methods Workbooks that explain the methodology used to prepare Australia’s
national GHG inventories (AGO 1998, 1999, 2001a). These emission factors are 4.2 Mg
CH4 /PJ (GCV) and 4.1 Mg N2O / PJ (GCV) (AGO 1998, 1999).

3.0 CHILE – FIRST NATIONAL COMMUNICATION TO THE CONFERENCE


OF THE PARTIES OF THE CONVENTION, OCTOBER 1999 (Chilean
National Environmental Commission 1999)

3.1 Methane and Nitrous Oxide

The report indicates that for non-CO2 gases, the Final Use Approach and default emission
factors are used, as recommended by the IPCC. Thus, it appears that IPCC’s Tier 2 method
with default emission factors was used for methane and nitrous oxide. The report does not
list the emission factors used. In the summary tables, there are no methane or nitrous oxide
emissions listed for the pulp and paper sector (Chilean National Environmental Commission
1999).

4.0 FINLAND – ANNEX B OF “GREENHOUSE GAS EMISSIONS AND


REMOVALS IN FINLAND” (Technical Research Center of Finland 2001)

4.1 Methane

Finland’s report to UNFCCC indicates that “CH4 and N2O emission factors are based on
mainly national research (included evaluation and application of international research and
measurement data)… IPCC default values are used to a lesser degree.” The emission factors
given in the Annexes to the report, however, are only for small units (less than 1 MW)
(Technical Research Center of Finland 2001).

C-2
4.2 Nitrous Oxide

Finland’s report to UNFCCC indicates that “CH4 and N2O emission factors are based on
mainly national research (included evaluation and application of international research and
measurement data)… IPCC default values are used to a lesser degree.” The emission factors
given in the Annexes to the report, however, are only for small units (less than 1 MW)
(Technical Research Center of Finland 2001).

The Finland report also indicates that “indirect emissions due to atmospheric deposition of
NH3 and NOX emissions…are calculated…for energy-related and industrial NOX emissions
using the IPCC methodology and default emission factor (0.01 kg N2O-N/kg NOX-N
emitted). The indirect N2O emissions caused by nitrogen deposition due to NOX emissions
in the energy sector are included in the emission estimates for the relevant sectors.” This
does not appear to be required by the IPCC estimation methods for energy production,
although it is mentioned in the IPCC’s methods for estimating GHG releases from agriculture
(Technical Research Center of Finland 2001). (See Section 4.5.4.A and Table 4-23 in the
IPCC Reference Manual.)

5.0 CANADA – “GREENHOUSE GAS ACTION PLAN GUIDELINES – 2001”


(FPAC 2001) AND “CANADA’S GREENHOUSE GAS INVENTORY, 1997
EMISSIONS AND REMOVALS WITH TRENDS” (Environment Canada 1999)

5.1 Methane

The method for estimating GHG emissions from biomass fuel combustion is generally
comparable to IPCC’s Tier 1 sectoral approach except that FPAC (and Environment Canada)
biomass emission factors are based on mass of fuel rather than heat content. The emission
factors are published in Appendix A of the FPAC Guidelines. The emission factor for
methane from wood fuel/wood waste is 0.15 g CH4/kg fuel (Environment Canada 1999 and
FPAC 2001).

In “Canada’s Greenhouse Gas Inventory, 1997 Emissions and Removals with Trends,”
Appendix B1 explains that the emission factors were derived as averages of those available
in EPA’s AP-42. The AP-42 factors have since changed. See the discussion of the US
protocol below for the most recent AP-42 emission factors. There is no emission factor
given for spent pulping liquors. Although the guidelines do not indicate whether the factors
are based on wet or dry weight of fuel, the emission factor shown for CO2 from wood
fuel/wood waste (1500 g/kg) suggests that the factors are based on dry weight. There is no
discussion of the effectiveness of control devices, but it is assumed that the factors are for
emissions after control devices. This is an unimportant departure from IPCC methods since
IPCC assumes that there is no removal of methane across control devices.

C-3
5.2 Nitrous Oxide

Like methane, methods for estimating N2O emissions from biomass fuel combustion are
comparable to IPCC’s Tier 1 methods except that (a) the emission factors are mass- or
volume-based rather then energy-based, and (b) there is no discussion of a correction for the
effectiveness of control devices (an unimportant difference since IPCC provides removal
data only for gas fired boilers.equipped with selective catalytic reduction). The emission
factor for nitrous oxide from wood fuel/wood waste is .016 g N2O/kg fuel. It is assumed that
this emission factor estimates emissions based on dry weight of fuel and after control devices
(Environment Canada 1999 and FPAC 2001).

In “Canada’s Greenhouse Gas Inventory, 1997 Emissions and Removals with Trends,”
Appendix B1 explains that the emission factors were derived as averages of those available
in EPA’s AP-42, which have since changed. See the discussion of the US protocol below for
the most recent AP-42 emission factors.

6.0 JAPAN – INFORMATION FROM THE JAPAN PAPER ASSOCIATION


(JPA 2001)

6.1 Methane and Nitrous Oxide

Methane and nitrous oxide are not included in the estimates collected by the Japan Paper
Association for the Federation of Economic Organization (Keidanren). The Japan Ministry
of Environment, however, includes CH4 and N2O in its inventories and uses the following
emission factors for biomass fuels (JPA 2001).

Table 1. Japan Ministry of the Environment Emission Factors for Biomass Fuels
from information provided by the Japan Paper Association (JPA 2001)
Combustion Unit Fuel CH4 N2O
Boiler Black Liquor 0.000033 kg/kg dry solids -
Boiler Wood 0.00082 kg/kg dry solids 0.000063 kg/kg dry solids
Waste combustion Sludge from mills 0.0097 kg/wet ton 0.45 kg/wet ton
at 50% moisture

7.0 SWEDEN – SWEDEN’S NATIONAL INVENTORY REPORT (Swedish


Environmental Protection Agency 2001)

7.1 Methane

Sweden uses an IPCC Tier 1 approach for estimating methane emissions. The emission
factors are contained in Table 1.7.3 of its 2001 Inventory Report. The emission factors for
methane from wood waste and black liquor combustion are both 30 kg CH4/TJ, the same as
IPCC’s Tier 1 default factor (Swedish Environmental Protection Agency 2001).

C-4
7.2 Nitrous Oxide

Like CO2 and methane, Sweden uses an IPCC Tier 1 approach for estimating nitrous oxide
emissions from fossil fuel fired units producing steam and/or power. The emission factors
are in Table 1.7.4 of its 2001 inventory. The Swedish emission factors for nitrous oxide from
wood waste and black liquor combustion are both 5 kg N2O/TJ, whereas IPCC’s Tier 1
default factor is 4 kg N2O/TJ. The factors are not identified as being uncontrolled, so it is
assumed that they reflect actual (i.e., controlled) emissions (Swedish Environmental
Protection Agency 2001).

8.0 UNITED KINGDOM – UK GREENHOUSE GAS INVENTORY, 1990 – 1999


(AEA Technology 2001)

8.1 Methane

The UK approach differs from the IPCC Tier 1 approach primarily in that the emission
factors for wood fuels are based on mass of fuel rather than energy content. The emission
factors used by the UK in developing its inventory are included in Tables 1 to 4 of Appendix
2 to “UK Greenhouse Gas Inventory, 1990-1999.” The emission factor for methane from
wood as a fuel in industry is 0.482 kg CH4/t. Although the table does not indicate whether
this is based on wet or dry tons, the emission factor for carbon from wood burning (470 kg/t)
suggests that the emission factors are based on dry weight. It is assumed that this factor is
for emissions after control devices have been applied (AEA Technology 2001).

8.2 Nitrous Oxide

The UK approach differs from the IPCC Tier 1 approach primarily in that the emission
factors for wood fuels are based on mass of fuel rather than energy content. The emission
factors used by the UK in developing its inventory are included in Tables 1 to 4 of Appendix
2 to “UK Greenhouse Gas Inventory, 1990-1999.” The emission factor for nitrous oxide
from wood as a fuel in industry is 0.069 kg N2O/t. Although the table does not indicate
whether this is based on wet or dry tons, the emission factor for carbon from wood burning
(470 kg/t) suggests that the emission factors are based on dry weight. It appears that the UK
emission factors represent final emissions (i.e., following control technologies) (AEA
Technology 2001).

C-5
9.0 UNITED STATES – INVENTORY OF U.S. GREENHOUSE GAS EMISSIONS
AND SINKS: 1990-1999 (USEPA 2001a) and AP-42 EMISSIONS FACTORS
(USEPA 2001b)

9.1 Methane

In its GHG inventory, EPA used IPCCs Tier 1 default methodology and emission factors to
estimate methane emissions from biomass fuels. The only emission factor is for wood – 30
kg CH4/TJ (USEPA 2001a).

USEPA’s AP-42 Emission Factors Document (Supplement G, July 2001) indicates that
“[m]ethane emissions are highest during periods of low-temperature combustion or
incomplete combustion, such as the start-up or shut-down cycle for boilers. Typically,
conditions that favor formation of N2O also favor emissions of CH4” (USEPA 2001). The
emission factor given in Supplement G as the average emission factor for methane from
wood residue combustion is 0.021 lb/MMBtu GCV (9.5 kg CH4/TJ NCV assuming that NCV
is 5% less than GCV). There are no data on non-CO2 GHG emissions from spent liquor
combustion in AP-42 (USEPA 2001b).

9.2 Nitrous Oxide

In its GHG inventory, EPA used IPCCs Tier 1 default methodology and emission factors to
estimate nitrous oxide emissions from biomass fuels. The only emission factor is for wood –
4 kg N2O/ TJ (USEPA 2001a).

USEPA’s AP-42 Emission Factors Document (Supplement G, July 2001) indicates that
“[f]ormation of N2O during the combustion process is governed by a complex series of
reactions and its formation is dependent upon many factors. Formation of N2O is minimized
when combustion temperatures are kept high (above 1475oF) and excess air is kept to a
minimum (less than 1 percent)” (USEPA 2001b). In earlier supplements to AP-42, EPA had
given separate emission factors for stoker boilers and fluidized boilers based on a theory that
fluidized bed boilers generated more N2O. In the current supplement, this distinction has
been eliminated and a single N2O emission factor is shown for all wood fired boilers. The
emission factor given in Supplement G as the average emission factor for nitrous oxide from
wood residue combustion is 0.013 lb/MMBtu GCV (5.9kg N2O/TJ NCV assuming that NCV
is 5% less than GCV). There are no data on non-CO2 GHG emissions from spent liquor
combustion in AP-42 (USEPA 2001b)

10.0 WRI/WBCSD – THE GREENHOUSE GAS PROTOCOL AND SUPPORTING


GUIDANCE (WRI 2001a, 2001b, 2001c)

10.1 Methane and Nitrous Oxide

The WRI/WBCSD guidance does not include CH4 and N2O. In the words of the October
2001 guidance, “a much greater effort is required to estimate CH4 and N2O emissions from

C-6
stationary sources [compared to CO2], and a much higher level of uncertainty exists.
Consequently, due the relative insignificance of CH4 and N2O emissions from most
stationary sources, [the WRI/WBCSD protocol] only includes guidance for estimating CO2
emissions from stationary sources” (WRI 2001b)

11.0 OTHER DATA ON NON-CO2 GHG EMISSIONS FROM BIOMASS


COMBUSTION

In reviewing the emission factors most commonly used for methane and nitrous oxide from
biomass combustion, one finds that most are based on very sparse data. In several cases,
protocols suggest using factors from the US EPA that are no longer supported by that agency.

Fortunately, in the last few years, additional data have been generated that, while not widely
known, provide an improved basis for estimating these emissions.

11.1 Finnish Data from Fortum Power and Heat Oy (Fortum 2001)

Perhaps the most ambitious recent study was undertaken in Finland by Fortum Power and
Heat Oy with funding from the Finnish Forest Industries Federation and the Finnish Energy
Industries Federation. The report “Methane and Nitrous Oxide Emissions in the Finnish
Energy Production” was published in May of 2001 and contains new data from 11 boilers
and 2 recovery furnaces (Fortum 2001). Of the 11 boilers, six were fluidized bed boilers
burning biomass and/or peat, two were pulverized coal-fired boilers, two were combined
cycle gas turbine facilities, and one was a heavy oil boiler. In addition, the report
summarizes existing emission factors, noting that most of them had been proven to be
unreliable.

The results of the study are summarized and contrasted to earlier emission factors in a table
reproduced below. The authors conclude that “fixed emission factors do not work well in
CFB boilers. The emission factor has to be such that it considers both the boiler load and
fuel” (Fortum 2001).

Table 2. Comparison of measured N2O and CH4 emissions with other emission factors.
-Parenthesis indicate abnormal process conditions-
Table 14 from Fortum Power and Heat Oy 2001
CFB- BFB-boilers PC- Gas turbine Recovery Heavy oil
boilers boilers plants boilers boiler
N2O national factor mg/MJ 30 - 39 27 - 30 2 1 1.4 2
N2O IPCC- factor mg/MJ 4 4 1.4 0.1 4 0.6
N2O measured mg/MJ <2...26 <2 <1 <3 <1 <1
CH4 national factor mg/MJ 11.1 - 22.4 14.5 - 21.3 4 3 1 8
CH4 IPCC- factor mg/MJ 30 30 1 1 30 3
CH4 measured mg/MJ 1 1 - 3 (72) <1 <1 (26) 1 <1

C-7
Using data from the Fortum report, NCASI has estimated that the median emission factors
for circulating fluidized bed boilers burning peat or bark were 1 kg CH4/TJ and 8.8 kg
N2O/TJ. The median emission factors for bubbling fluidized bed boilers burning peat or bark
were 2 kg CH4/TJ (excluding one very large number associated with low oxygen-high carbon
monoxide conditions) and <2 kg N2O/TJ.

11.2 US Data from NCASI

Although not as recent or extensive, NCASI developed methane data in the late 1970s and
early 1980s (NCASI 1980, 1985). The results are shown in the following table. The
methane emissions fall within the range of the newer Finnish data, but tend to be a little
higher, on average. This is perhaps not surprising since the NCASI data were generated on
older technology boilers during a period when there was not as much emphasis on
combustion control. Still, the data are fairly consistent, and in any event, far lower than
IPCC’s default value.

Table 3. Data on Methane Emissions from Wood Residue Boilers


Collected in the Late 1970s and Early 1980s Generated by EPA Method 25
Data have been converted from GCV to NCV assuming a 5% difference
Reference Boiler Types Number Methane Emissions
of mg/MJ NCV
Boilers
Min. Max. Ave.
NCASI Air Technical 1952 to 1976 vintage, stoker- 4 <1 20.4 8.2
Bulletin 109, Sept. fed boilers, sampled before
1980 (NCASI 1980) particulate control devices
NCASI Technical 1952 to 1981 vintage, stoker- 6 <1 4.5 2.7
Bulletin 455, April fed boilers, all but one
1985 (NCASI 1985) sampled after wet scrubbers

C-8
ANNEX D

GREENHOUSE GAS EMISSIONS ATTRIBUTABLE TO IMPORTS AND EXPORTS


OF POWER AND STEAM – OVERVIEW OF METHODS IN EXISTING
PROTOCOLS

1.0 IPCC – REVISED 1996 IPCC GUIDELINES FOR NATIONAL


GREENHOUSE GAS INVENTORIES AND MAY 2000 GOOD PRACTICES
DOCUMENT (IPCC 1997a, 1997b, 1997c, 2000)

Because the IPCC guidelines are for national inventories, they have very little guidance on
how to apportion emissions related to facility-level imports and exports of power. The most
relevant discussion regards methods for developing sectoral level emissions estimates. The
discussion below is based on the 1996 Revised IPCC Guidelines for sectoral level reporting.

In Volume 1 – Reporting Instructions for IPCC’s Revised 1996 guidance, Chapter 1


describes the general method used by IPCC to assign emissions to various sectors (IPCC
1997a). Energy-related emissions from the “Pulp, Paper, and Print” industry (category
1.A.2.d) are included in the “Manufacturing Industries and Construction” (category 1.A.2)
portion of the reporting framework. The guidance explains that this part of the framework is
to include “[e]missions from combustion of fuels in industry including combustion for the
generation of electricity and heat. Emissions from autoproducers [companies that produce
their own electrical power] should be assigned to the sector where they were generated and
an attempt made to separately identify the emissions associated with autogeneration from
those associated with process heat” (IPCC 1997a, page 1.4).

In this same section, autoproducers are defined as “undertakings which generate


electricity/heat wholly or partly for their own use, as an activity which supports their primary
activity.” In Volume 3, the Reference Manual, there is additional elaboration in Box 3 on
page 1.32 (IPCC 1997a). “An autoproducer of electricity or heat is an enterprise which
generates electricity or sells heat as a secondary activity, i.e., not as its main business. This
should be contrasted with the main power producers who generate electricity or who sell heat
as their main business (primary activity) and may be publicly or privately owned. Supplies
from main power producers are referred to as “Public” electricity and heat supply although
an increasing part of public supply is being met by autoproducers” (IPCC 1997a)

The section on “Generation of electricity and heat” in the Reference Manual indicates that
“emissions from autoproduction are attributed to the industrial or commercial branches in
which the generation activity occurs. Emissions reported under ‘Public Electricity and Heat’
should be those from main power producers only” (IPCC 1997c, pg 1.31 and 1.32).

Thus, it is clear that IPCC’s method places all energy-related emissions from pulp and paper
mills in the pulp and paper sector unless the emissions are from an activity where pulp and
paper making is not the primary business. The emissions from boilers making steam, some
of which is exported or used to make exported electricity, remain in the pulp and paper sector

D-1
unless the boilers are associated with a separate business that just happens to be located on
the mill site. Similarly, the emissions from gas turbines on mill sites are attributed to the
pulp and paper sector unless, perhaps, the turbines are owned by a separate company whose
main business is power generation.

Likewise, emissions from the off-site production of electricity used by pulp and paper mills
remain in the energy sector rather than being assigned to the pulp and paper sector.

It is clear from all of this that IPCC’s objective is to eliminate double counting in national
inventories by providing guidance as to how energy-related emissions should be reported.
IPCC’s general intention is that emissions be reported in the sector where they were
generated and IPCC puts a facility, and all of its emissions, into the sector representing its
primary activity.

2.0 AUSTRALIA – NATIONAL GREENHOUSE GAS INVENTORY – 1999


(Australian Greenhouse Office 1998, 1999, 2001a, 2001b and PPMFA 2001)

The Pulp and Paper Manufacturers Federation of Australia collects greenhouse gas emissions
information in an annual Environmental Performance Indicators survey (PPMFA 2001). The
survey form indicates that emission factors for electricity (as well as gas and coal) vary
between states. Companies are directed to a Factors and Methodologies Workbook that is
published by the government under its Greenhouse Challenge program.

The Factors and methodologies Workbook includes purchased power emission factors for
each State. The emission factors include transmission and distribution losses (AGO 2001b).
The emission factors are shown below.

Table 1. Australian Emission Factors for Electricity Purchased (AGO 2001b)


State kg CO2-equiv./kWh*
New South Wales, Australian Capital Territory 0.968
Victoria 1.467
Queensland 1.040
South Australia 1.109
Western Australia 1.032
Tasmania 0.002
Northern Territory 0.756
* Includes transmission and distribution losses

D-2
3.0 FINLAND – ANNEX B OF “GREENHOUSE GAS EMISSIONS AND
REMOVALS IN FINLAND” (Technical Research Center of Finland 2001)

Finland’s national inventory follows the IPCC practice of allocating emissions to the sector
where they are generated. There is no attempt, therefore, to allocate emissions from power
producers to industrial consumers of purchased power.

The large amounts of combined heat and power (CHP) production in Finland, however, have
prompted that country to use a slightly different approach when deciding in which sector to
place power plants that supply a large fraction of their power or steam output to industrial
facilities. Annex B indicates that “[i]n Finland many CHP power plants produce electricity
and steam to the manufacturing industry, especially to pulp and paper plants. These power
plants are usually economically part of the industrial sectors themselves, and in this case, the
power plants are regarded as 'autoproducers' according to the IEA definitions. In the Finnish
inventory (CRF tables) these autoproducer power plants are allocated to the corresponding
industrial sectors. During the last years a growing number of these autoproducer power plants
have been sold to the energy companies (i.e., companies, the primary activity of which is to
produce and sell electricity and heat to the market). By definition, these plants should be
allocated to the Energy Industries sector [under the IPCC protocol]. However, in the Finnish
inventories these plants are treated as autoproducers. Also some new power plants which
have been built by the energy companies to serve the manufacturing industry have been
categorized in the same way in the Finnish inventory. This allocation has been used in cases
where there is a direct connection (steam pipe) between the power plant and the industrial
plant and the industrial plant uses most of the energy (usually steam) produced by the power
plant. The reasoning and justification for this allocation is that it enables linking of the
emissions more closely to the actual production processes” (Technical Research Center of
Finland 2001).

The Finnish Forest Industries Federation periodically publishes environmental statistics for
the Finnish industry (Finnish Forest Industries Federation 2001). This document indicates
that the forest industry’s electricity requirement in 2000 was 26.3 billion kilowatt-hours,
corresponding to 61% of industry’s and 33% of Finland’s total requirement. Of the 26.3
billion kWh required by the industry, about 11.1 billion kWh was generated “in-house.” The
amounts procured from outside were largely from energy producers in which forest industry
companies are shareholders. The report contains separate statistics for emissions from
purchased power. No details are given in the report regarding how these indirect emissions
are calculated for power or steam exports, but the Finnish Forest Industries Federation has
provided the following information.

The Finnish Forest Industries Federation asks mills to report only direct emissions. These
are the same reports as made to the Finnish authorities. Where estimates of indirect
emissions are needed, the Finnish Forest Industries Federation estimates the indirect
emissions by using mill data on the amounts of purchased electricity together with
information on the fuel mix of the electrical power generation plants supplying the mills. It
is unclear what is done in cases where the mill is purchasing steam (instead of, or in addition

D-3
to, electricity) from an off-site entity. In cases where mills export electricity, no attempt is
made to reduce the mill’s reported emissions to reflect the exports. (Karessuo 2001).

4.0 CANADA – “GREENHOUSE GAS ACTION PLAN GUIDELINES – 2001”


(FPAC 2001) AND “CANADA’S GREENHOUSE GAS INVENTORY, 1997
EMISSIONS AND REMOVALS WITH TRENDS” (Environment Canada 1999)

FPAC recommends that companies “report indirect emissions that they influence as part of
their production process” (FPAC 2001). These are reported separately from direct emissions.
Emission factors are supplied to convert purchased electricity into CO2 emissions. The
factors vary Province by Province reflecting the methods and fuels used to generate
electricity. These factors, which are also published in the 1999 Voluntary Challenge and
Registry (VCR) Guidelines are shown in the following table (VCR 1999).

The FPAC guidelines do not directly address methods for estimating the impacts of electrical
power exports from industry facilities.

The special case of power imports or exports from CHP production are not addressed directly
in the FPAC guidelines, but the guidelines point out that under the Canadian Baseline
Protection Initiative, “parties involved in cogeneration projects may decide between
themselves how to share direct reductions” (FPAC 2001)

Table 2. Electric Energy Consumption Indirect Emissions Factors


(Table 5.4 from VCR 1999)
Province kg CO2 equivalents / kWh
Newfoundland and Labrador 0.19
Prince Edward Island 0.546
Nova Scotia 0.78
New Brunswick 0.546
Quebec 0.0014
Ontario 0.18
Manitoba 0.011
Saskatchewan 0.862
Alberta 0.991
British Columbia 0.02
Yukon/Northwest Territories/ Nunavut GHG emission rates depend on
local generation equipment.

D-4
5.0 FRANCE – “INVENTAIRE DES ÉMISSIONS DE GAZ À EFFET DE SERRE
EN FRANCE DE 1990 À 1999” [INVENTORY OF GREENHOUSE GAS
EMISSIONS IN FRANCE FOR THE PERIOD 1990-1999] (CITEPA 2000)

The French national inventory includes a sectoral level analysis, built on the CORINAIR
methodology. The report indicates that emissions are assigned to the sector where they are
emitted. As an example, the report indicates that there is no emission penalty assigned for
using electricity because the emissions are assigned to the entity producing the electricity
(CITEPA 2000, Appendix 3, page 13).

The report goes on to say that at facilities that generate their own electricity, the emissions
are assigned to the sector representing the principal activity of the facility (e.g., industrial,
district heating, etc.) (CITEPA 2000, Appendix 3, page 15).

6.0 JAPAN – INFORMATION FROM THE JAPAN PAPER ASSOCIATION (JPA


2001)

With regards to electricity or steam that is purchased or sold by mills, the Japan Paper
Association (JPA) either adds or deducts energy used and/or carbon emissions in the
following manner.

For purchased electricity, JPA uses emission factors published by the national organization
representing the electrical power producers. For the year 2000, the factors were –

1 kWh = 9.00 MJ
kWh = 0.090 kg Carbon
(Note: The value used by the Ministry of Environment for the pulp and paper
industry in 1999 was 0.097 kg Carbon/kWh or 0.357 kg CO2/kWh)

For purchased steam, the factor is 1 kg steam = 2,675 kJ in fuel. This would be converted
into CO2 emissions by considering the carbon content of the fuel. (Note: The value used by
the Ministry of Environment for the pulp and paper industry in 1999 was 0.072 kg CO2 / MJ
steam.)

For electricity sold, mills calculate the emissions associated with the sold power by
considering the fuels used to generate the electricity, their carbon contents, and the efficiency
of electrical power production. Sold steam is considered in the same way except boiler
efficiency is used (rather than the overall efficiency of power production) (JPA 2001).

Where it is necessary to apportion emissions because energy is purchased or sold in more


than one form, the emissions are allocated by considering the energy content of the steam or
electricity being sold (or purchased) and the overall efficiency associated with its generation.
From these values, one can estimate the amount of fuel that was required to generate the
specific amount of steam or electricity. The carbon emissions are then calculated from the
carbon contents of the fuel (JPA 2001).

D-5
7.0 SWEDEN – SWEDEN’S NATIONAL INVENTORY REPORT 2001 AND
INFORMATION FROM THE SWEDISH FOREST INDUSTRIES
FEDERATION

Sweden’s report contains methods for the national inventory only, and does not discuss the
issue of direct versus indirect emissions (Swedish Environmental Protection Agency 2001).

The Swedish Forest Industries Federation reports GHG emissions from the Swedish industry.
It estimates indirect emissions based on reported purchases of electricity and emission
factors. In cases where a mill sells (or gives away) power or steam, it retains the emissions
associated with its generation (Axelsson 2001).

8.0 UNITED KINGDOM – UK GREENHOUSE GAS INVENTORY, 1990 – 1999


(AEA Technology 2001) AND INFORMATION FROM THE PAPER
FEDERATION OF GREAT BRITAIN (Barnard 2001)

The UK constructs its inventory from data generated in a number of national programs that
must be aligned with the IPCC methodology. Among the adjustments that must be made is
the method for handling emissions from autogeneration of electrical power in CHP systems.
In the UK inventory, “CHP systems where the electricity is fed into the public supply are
classified as power stations and CHP systems where the electricity is used by the generator
are classified as autogeneration.” CHP autogenerator emissions are included in the “other
industry” sector of the sectoral energy tables in the IPCC Common Reporting Framework
(i.e., sector 1A2f) (AIA Technology 2001).

Otherwise, the UK report follows the IPCC convention of assigning the emissions to the
sector where they originate. The issue of direct versus indirect emissions is not otherwise
dealt with.

The Paper Federation of Great Britain indicates that because its agreement with the
government regarding the Climate Change Levy is based on energy efficiency, the question
of indirect emissions has not been addressed in this context (Barnard 2001).

9.0 UNITED STATES – INVENTORY OF U.S. GREENHOUSE GAS EMISSIONS


AND SINKS: 1990-1999 (USEPA 2001a) and GUIDELINES FOR THE
DEPARTMENT OF ENERGY 1605(B) PROGRAM (USDOE 2000)

Although EPA’s US inventory is not intended to allocate emissions at the company- or


facility-level, it does attempt to divide national GHG emissions among different sectors of
the economy, estimating “combustion” and “electricity” emissions from the industrial,
transportation, residential, and commercial sectors. The emissions generated by electrical
power producers are “distributed to each end-use sector based upon the sector’s share of

D-6
national electricity. … This method of distributing emissions assumes that each sector
consumes electricity generated from an equally carbon-intensive mix of fuels and other
energy sources. In reality, sources of electricity vary widely in carbon intensity (e.g., coal
versus wind power). By giving equal carbon-intensity weight to each sector's electricity
consumption, emissions attributed to one end-use sector may be somewhat overestimated,
while emissions attributed to another end-use sector may be slightly underestimated.”
(USEPA 2001a, pg. 2-7)

This method misses the amounts of power generated by non-utility generators. In EPA’s
words, “[a]ccording to current EIA [the Energy Information Administration of the
Department of Energy] sectoral definitions, the industrial sector also includes emissions from
nonutility generators (e.g., independent power producers) who produce electricity for their
own use, to sell to large consumers, or to sell on the wholesale electricity market. The
number of nonutility generators and the quantity of electricity they produce has increased
significantly as many States have begun opening their electricity markets to competition. In
future inventories, these nonutility generators will be removed from the industrial sector and
incorporated into a single electric power sector with electric utilities.” (USEPA 2001a, pg. 2-
10)

At present, therefore, EPA assigns to the industrial sector all emissions generated by fossil
fuel consumption at industrial facilities and adds to this, the emissions generated in the
production of power consumed by the industrial sector. For each sector, direct and indirect
emissions are reported separately in the sector-level reporting by EPA. Indirect emissions
are calculated using the average emission factor for the national electrical power grid. In the
future, EPA will attempt to remove nonutility electricity generators from the industrial sector.
Many of these companies have as their primary business, electrical power generation. It is
not known how EPA will allocate emissions from power generators that produce electricity
primarily for themselves, but may export excess power or heat (USEPA 2001a).

The Department of Energy has a different set of guidelines that are focused on facility- or
company-level reporting. This different focus is due to the fact that the DOE guidelines are
for use in a program for reporting voluntary reductions in GHGs at the facility- or company-
level (the 1605(b) program). Generally, the guidelines are very flexible because DOE wants
to encourage a wide range of GHG reduction activities, some of which may not lend
themselves to standard carbon accounting methods. The guidelines specifically allow the
reporting of reductions in either direct or indirect emissions, or both (USDOE 2000).

DOE provides guidance for estimating emissions related to purchased power (or reduced use
of purchased power). It suggests using emission factors developed by DOE for average
electrical power generated in each of the 50 states or case-by-case emission factors derived
from information on the fuels used to generate the purchased power. The DOE last updated
the state-specific emission factors for electricity in 2000. The factors, and the methodology
to estimate them are in “Updated State-level Greenhouse Gas Emission Factors for
Electricity Generation” published in March 2001 by the Energy Information Administration
(USDOE 2001a). The emission factors from that document are available on the internet and
are shown in the following table (USDOE 2001b).

D-7
Table 3. 1997-99 US State Average CO2 Emission Factors for Electric Utilities
State t CO2/MWh State t CO2/MWh
Alabama 0.611 Montana 0.586
Alaska 0.563 Nebraska 0.638
Arizona 0.461 Nevada 0.755
Arkansas 0.594 New Hampshire 0.338
California 0.138 New Jersey 0.257
Colorado 0.929 New Mexico 0.952
Connecticut 0.515 New York 0.361
Delaware 0.842 North Carolina 0.564
Florida 0.65 North Dakota 0.993
Georgia 0.644 Ohio 0.823
Hawaii 0.824 Oklahoma 0.802
Idaho 0 Oregon 0.08
Illinois 0.534 Pennsylvania 0.554
Indiana 0.968 Rhode Island 0.418
Iowa 0.899 South Carolina 0.37
Kansas 0.79 South Dakota 0.325
Kentucky 0.901 Tennessee 0.565
Louisiana 0.603 Texas 0.702
Maine 0.408 Utah 0.88
Maryland incl. D.C. 0.613 Vermont 0.008
Massachusetts 0.561 Virginia 0.493
Michigan 0.79 Washington 0.083
Minnesota 0.72 West Virginia 0.895
Mississippi 0.599 Wisconsin 0.813
Missouri 0.862 Wyoming 0.995

The USEPA also provides emission factors for purchased electrical power in its Emissions &
Generation Resource Integrated Database (E-GRID). The information in E-GRID is based
on an integration of multiple federal data sources and includes information on non-utility
power generators as well as utility generators. The E-GRID information can be downloaded
from the internet as a self-contained, stand-alone “data browser” program and as a series of
spreadsheets (available at http://www.epa.gov/airmarkets/egrid/index.html).

10.0 WRI/WBCSD – THE GREENHOUSE GAS PROTOCOL AND SUPPORTING


GUIDANCE (WRI 2001a, 2001b, 2001c)

The WRI/WBCSD Greenhouse Gas Protocol identifies three “scopes” for emissions
reporting. Scope one is direct emissions, scope two is indirect emissions from imports of
electricity, steam or other heat, and scope three is other indirect emissions. The GHG
Protocol recommends that companies account for and report at least scopes 1 and 2.

D-8
Emissions associated with the export of electricity or steam are included as “supporting
information” and not netted against a facility’s direct emissions. Only CO2 must be reported
(WRI 2001a, 2001b).

The October 2001 Greenhouse Gas Protocol provides the following additional guidance for
reporting indirect emissions from imported electricity, steam and other heat (WRI 2001a).

“• Emissions from imported electricity can be estimated from purchase records and
grid emissions factors. You should use the most reliable emissions factors available
and be consistent in their use.
• Purchases of electricity by electric utilities for sale to end-use customers (e.g., an
electricity utility that has a supply contract with a power generator) should be
reported under scope 2. The rationale for this is that utilities often exercise choice
over where they buy their energy and this may present significant opportunities for
GHG reductions.
• Trading transactions of electricity should not be reported.
• If you export electricity, heat, or steam to the grid or to another company, the
emissions associated with the exports should not be deducted from scope 1.
• Emissions from exported electricity, heat, or steam should be reported under
supporting information and not deducted from any imports, as this would be
inconsistent with how other exported products are accounted, e.g., export of clinker
by a cement company or scrap steel by an iron and steel company.
• GHG emissions from activities upstream of your electricity provider, e.g.,
exploration, drilling, flaring, transportation, and refining should not be reported under
scope 2.”

The WRI/WBCSD protocol gives three options for selecting emission factors for imports of
electricity or steam. Listed from most preferred to least preferred, these are (a) an emission
factor supplied by the generator of the electricity or steam based on the fuels used or the
purchasing agreement (i.e., for “green” power), (b) emission factors published by authorities
for the locality or region, and (c) national average emission factors (WRI 2001b).

The calculation tools provided for the GHG Protocol contain emission factors that do not
incorporate transmission and distribution losses (WRI 2001b, 2001c). WRI/WBCSD
indicates that these losses can vary from 5% in OECD countries and up to 20% in developing
countries (WRI 2001b).

The WRI/WBCSD worksheet “Calculating CO2 emissions from stationary combustion”


contains a number of spreadsheets, including several of interest in the context of estimating
indirect emissions (WRI 2001c). They deal with
· emissions resulting from electricity and/or steam import (worksheet 2),
· emissions resulting from electricity and/or steam export to grid or through direct supply
contract (worksheet 3),
· emissions from CHP systems calculated using the “efficiency of heat and power
production” (worksheet “CHP option 1”), and

D-9
· emissions from CHP systems calculated using the “work potential” method (worksheet
“CHP option 2”).

Emissions attributed to imported electricity or steam are calculated in worksheet 2 simply by


multiplying the appropriate emission factor by the amounts or power or steam imported.

In worksheet 3, the impact of electricity or steam exports is calculated by dividing the


facility’s electricity and steam generation-related emissions between facility emissions and
exported emissions according to the relative amounts of electricity/steam used internally and
exported.

Worksheets “CHP Option 1” and “CHP Option 2” are of interest in the context of indirect
emissions because they attempt to divide the emissions from CHP systems between the steam
and power produced, a topic addressed in detail in Annex E of these tools.

D - 10
ANNEX E

ALLOCATING GREENHOUSE GAS EMISSIONS FROM COMBINED


HEAT AND POWER (CHP) SYSTEMS – RECOMMENDED GUIDANCE
AND REVIEW OF METHODS

This Annex contains the material presented on the efficiency method in Section 12.7 of
Calculation Tools for Estimating Greenhouse Gas Emissions from Pulp and Paper Mills and
additional information on other methods for allocating CHP emissions.

1.0 RECOMMENDED GUIDANCE

Where electricity is produced by combined heat and power (CHP) systems, it may be
necessary to allocate the emissions from the CHP system to the various output energy
streams. Of course, if the mill owns the CHP system and uses all of its output, this is not
necessary because all the emissions are direct emissions for the mill. In many cases,
however, a mill may either receive CHP energy from an outside provider or export a portion
of its own CHP output. For instance, if a mill is importing steam from a nearby power plant,
it is necessary to estimate indirect emissions associated with the imported steam. Likewise,
if a mill is exporting power from a CHP system but using the steam internally, it may be
necessary to estimate how much of its emissions to attribute to the exported electricity.

Although there are several methods for allocating emissions from CHP systems, the
“efficiency” method is recommended in these calculation tools because it attempts to relate
energy outputs to the amounts of fuel used to generate them, and by extension, to the GHGs
produced in generating them. The efficiency method is one of two methods recommended
by WRI/WBCSD (termed “CHP option 1” by WRI/WBCSD) (WRI 2001b, c).

2.0 OVERVIEW OF METHODS

There are at least four methods that can be applied in a broad fashion for allocating GHG
emissions among electricity and steam or hot water outputs from CHP plants. All four
methods presented herein involve estimating total CHP system emissions based on fossil fuel
combustion and distributing the total emissions among the various output streams.
Allocations are made based either on the perceived value of the energy outputs, the “useful
energy” content of each energy output, or by estimating the amount of original fuel energy
expended in creating each energy output.

The “financial value” method of allocating emissions involves assigning a monetary value to
each energy output stream and allocating emissions according to the value of the energy.
The methods for determining these values are site specific, so no attempt will be made to
present alternative ways to use this allocation method. Therefore, the guidance recommends
that companies not use the financial method to allocate emissions from CHP systems.

The “efficiency” method is based on allocating emissions according to the amount of fuel
used to produce each energy output. The method uses either assumed or estimated

E-1
efficiencies for conversion of energy at various points in the process to back calculate the
amounts of fuel associated with each output energy stream. This method can be used in a
simplified or detailed manner, and is the approach recommended in the pulp and paper
industry calculation tools.

The “heat content” and “work potential” methods allocate emissions based on the amount of
useful energy in each energy output. Both of these allocation methods consider the energy
content of electrical power to be of “complete utility,” such that all of the energy in the
electricity is consumed in a useful fashion by a process. The primary difference between the
allocation methods is in regard to how the energy content associated with steam is
determined. The heat content method assumes that the useful energy content of steam (or hot
water) is equivalent to the heat that can be extracted from it, whereas the work potential
method assumes that the useful energy content is equivalent to the maximum amount of work
that can be extracted from the steam. Accordingly, the work potential method is not
recommended for allocating emissions from CHP systems which incorporate a hot water
energy output stream (work cannot be extracted from hot water).

The efficiency, heat content, and work potential methods will be briefly described below,
followed by illustrative examples of allocating GHG emissions for a hypothetical CHP
system by each method.

2.1 Efficiency Method – Note: Section 2.1.1 is identical to Section 12.7.1 in the
report

2.1.1 Simplified Efficiency Method

The efficiency method requires use of assumed efficiency factors for the production of power
and steam, or actual efficiency factors for each steam or power generation device based on
detailed process design and operating information. It is assumed that the efficiency of
producing hot water is the same as the efficiency of producing steam. The simplest approach
to applying the efficiency method is to assign a single efficiency factor to all power output
and a single efficiency factor to all heat (steam and hot water) output. This information is
used to compute an efficiency ratio equal to the heat production efficiency divided by the
power production efficiency. For example, if the CHP system produces steam at 80%
efficiency and power at 40% efficiency the ratio would be 2. The efficiency ratio is used
rather than the individual efficiencies because (a) it is the ratio that controls the allocation of
emissions rather than the individual efficiencies, and (b) the individual efficiencies are
constrained by the energy balance so it is not possible to specify both independently.
Emissions from the CHP system are allocated between the heat and power outputs, based on
this ratio of efficiencies, using Equations 1 and 2. This approach is referred to in this report
as the simplified efficiency method. The simplified efficiency method is the method
recommended for mills that lack, or choose not to use, detailed design and operating data
from CHP systems.

ìï H üï eH
EH = í ý ´ ET ; Reff = (Eq. 1)
ïî H + P ´ Reff ïþ eP

E-2
where: EH = emissions share attributable to heat production, t GHG/y
ET = total emissions from the CHP plant, t GHG/y
H = heat output, GJ/y
P = power output, GJ/y
Reff = ratio of heat production efficiency to power production efficiency
eH = assumed efficiency of typical heat production (default = 0.8)
eP = assumed efficiency of typical electric power production (default = 0.35)

The emission share attributable to electric power production is assigned from the relation:

E P = ET - E H (Eq. 2)

where: EP = emissions share attributable to electric power production

In these calculations, the heat in steam can be corrected to reflect the amount of heat in
returned condensates.

In using the simplified efficiency method, WRI/WBCSD recommends that an efficiency of


0.35 be used for power generation and 0.8 be used for steam (or hot water) generation (WRI
2001b, c), corresponding to a ratio of efficiencies (Reff) of 2.3. The first example calculation
below makes use of these recommended default efficiency factors.

Example Calculation: Allocating CHP emissions to three output streams – Simplified


efficiency method with WRI/WBCSD recommended default efficiency factors

A mill has the CHP system shown in the following figure, but it is lacking (or chooses not to
use) detailed energy balance information. Instead, the company chooses to use the simplified
efficiency method and the default efficiencies recommended by WRI/WBCSD; 0.35 for
power generation and 0.8 for steam generation (WRI 2001b, 2001c).

Fuel2
974 m3/hr nat. gas Steam P2=3 MW
HRSG
Stm. Turbine
Hs= 19.21 MW
Heat effB

H1=15 MW
Gas-fired Turbine 19,500 kg/hr steam
P1=5 MW 170°C, 7 bar
Fuel1
1538 m3/hr nat. gas

Using these assumed efficiencies, emissions can be allocated among the three outputs of the
CHP system as follows (using a basis of one hour of operation):

Total system emissions:

Fuel1:
CO2 (1538 m3/hr) x (0.039 GJ/m3) x (55.9 kg CO2/GJ) = 3353 kg CO2/hr

E-3
CH4 (1538 m3/hr) x (0.039 GJ/m3) x (0.0006 kg CH4 /GJ) x (21 CO2-eq. / CH4)
= 0.76 kg CO2-eq./hr
N2O (1538 m3/hr) x (0.039 GJ/m3) x (0.0001 kg N2O /GJ) x (310 CO2-eq. / N2O)
= 1.86 kg CO2-eq/hr
Total Fuel1 emissions = 3356 kg CO2-eq./hr

Fuel2:
CO2 (974 m3/hr) x (0.039 GJ/m3) x (55.9 kg CO2/GJ) = 2123 kg CO2/hr
CH4 (974 m3/hr) x (0.039 GJ/m3) x (0.0014 kg CH4 /GJ) x (21 CO2-eq. / CH4)
= 1.12 kg CO2-eq./hr
N2O (974 m3/hr) x (0.039 GJ/m3) x (0.0001 kg N2O /GJ) x (310 CO2-eq. / N2O)
= 1.18 kg CO2-eq./hr
Total Fuel2 emissions = 2126 kg CO2-eq./hr

Total CHP system emissions = 3356 + 2126 = 5482 kg CO2-eq./hr

Total system power output = P1 + P2 = 8 MW

0 .8
Reff = = 2.3
0.35

ì 15MW ü
EH = í ý ´ 5482kgCO 2 eq = 2462 kgCO 2 eq = 20,681 t CO2 eq/y at 350 d/y operation
î 15MW + (8 MW ´ 2.3) þ

E P = 5482 kgCO eq - 2452 kgCO eq = 3020 kgCO eq = 25,368 t CO2 eq/yr at 350 d/y operation
2 2 2

Using the simplified efficiency method with default power and steam efficiency factors,
therefore, the emissions from the CHP system are allocated to the output streams in the
following percentages:
· Percentage of CHP emissions to heat output = 100*2462/5482 = 44.9%
· Percentage of CHP emissions to power output = 100*3020/5482 = 55.1%

These percentages can be used to allocate all GHG emissions from the CHP system.
Emission factors can be developed for the energy outputs:
· Emission factor for CHP heat output = (2462 kg CO2-eq./hr)/15 MW
= 164.1 kg CO2-eq./MWh
· Emission factor for CHP power output = (3020 kg CO2-eq./hr)/8 MW
= 377.5 kg CO2-eq./MWh

2.1.2 Detailed Efficiency Method

Application of the relations in Equations 1 and 2 to allocate GHG emissions among the
energy outputs of a simple CHP system which includes only a single heat stream (in the form
of steam or hot water) and a single electric power stream may be fairly straightforward.
However, many industrial CHP systems include multiple heat output streams and incorporate
electric power production from multiple generators driven by different motive forces. To use

E-4
the efficiency method to allocate GHG emissions among the multiple energy outputs of more
complex CHP systems, Equations 1 and 2 can be modified to more general forms such as:

ì æ H1 ö ü
ï ç ÷ ï
çe ÷
ï è H1 ø ï
E H1 = í ý ´ ET (Eq. 3)
æ H ö æ H ö æ P ö æ P ö
ï ç 1 ÷ + ç 2 ÷ + ... + ç 1 ÷ + ç 2 ÷ + ... ï
ïç e ÷ ç e ÷ çe ÷ çe ÷ ï
î è H1 ø è H2 ø è P1 ø è P2 ø þ

ì æ P1 ö ü
ï çç ÷÷ ï
ï
EP1 = í
è eP1 ø ï
(Eq. 4)
ý ´ ET
ïç æ H1 ö æ H 2 ö æ P1 ö æ P2 ö
÷+ç ÷ + ... + çç ÷÷ + çç ÷÷ + ...ï
ïîçè eH1 ÷ø çè eH2 ÷ø è eP1 ø è eP2 ø ïþ

Where: EH1 = emissions share attributable to heat production as contained in steam


stream 1
EP1 = emissions share attributable to electric power production via generator 1
ET = total emissions from the CHP plant
H1 = heat output contained in steam stream 1
H2 = heat output contained in steam stream 2
P1 = power output from generator 1
P2 = power output from generator 2
eH1 = overall efficiency of producing heat contained in steam stream 1
eH2 = overall efficiency of producing heat contained in steam stream 2
eP1 = overall efficiency of producing electric power via generator 1
eP2 = overall efficiency of producing electric power via generator 2

Mills may already have energy balances that incorporate the type of information needed to
perform the detailed efficiency method. In these cases, the efficiency method is applied by
using the energy balances to estimate the amount of fuel required to produce each CHP
output stream. This can then be converted into GHG allocations for each stream. As in the
simplified efficiency method, hot water streams are treated in the same manner as steam
outputs.

E-5
Example Calculation: Allocating emissions from a complex CHP system.

Figure 1 depicts a hypothetical CHP system that includes three energy output streams (one
steam stream, H1, and two power outputs, P1 and P2) and incorporates two fuel inputs (one to
the gas-fired turbine and a second to the heat recovery steam generator (HRSG)). In order to
use Equations 3 and 4 to allocate GHG emissions among the three energy outputs of this
CHP system, efficiency factors for each output must be either developed or assumed.

Fuel2
974 m3/hr nat. gas
10.55 MW Steam P2=3 MW
HRSG
Stm. Turbine
Hs= 19.21 MW
19,500 kg/hr stm
Heat effB 540oC, 38 bar
10.83 MW

H1=15 MW
Gas-fired Turbine 19,500 kg/hr steam
P1=5 MW 170°C, 7 bar
Fuel1
1538 m3/hr nat. gas
16.67 MW

The CHP system emissions are the same as those calculated in the previous example:

Total Fuel1 emissions = 3356 kg CO2-eq./hr


Total Fuel2 emissions = 2126 kg CO2-eq./hr

The efficiency for P1, the power output from the gas-fired turbine, has been estimated at 0.3
(30%) based on information from the manufacturer. Mechanical losses in the gas turbine are
approximately 5%, so the “efficiency”1 of producing the (waste) heat in the turbine exhaust is
1-0.05-0.3=0.65, or 65%. The emissions from combustion of fuel in the gas-fired turbine can
now be allocated between P1 and the waste heat using Equations 3 and 4, with one hour of
operation as the basis for the calculations:

ì æ P1 ö ü
ï ç ÷ ï ì æ 5MW ö ü
çe ÷ ï ç ÷ ïï
ï è P1 ø ï ï è 0.3 ø
E P1 = í ý ´ E = í ý ´ 3356kgCO2 eq = 1678kgCO2 eq
ï æç P1 ö÷ æç Heat ö ï F1 ï æ 5MW ö æ 10.83MW ö ï
÷ ç ÷+ç ÷
ïç e ÷ + ç e ÷ï ï è 0.3 ø è 0.65 ø þï
î
î è P1 ø è Heat øþ

1
The term “efficiency” is used here to represent the amount of waste heat generated in the gas turbine relative
to the amount of fuel energy input to the gas turbine. Although waste heat generation rates are not typically
characterized by efficiency factors, the factor is required for the use of the efficiency method of emissions
allocation in this example because the waste heat from the gas turbine is an energy input to the HRSG.

E-6
ì æ Heat ö ü
ï ç ÷ ï ì æ 10.83MW ö ü
çe ÷ ï ç ÷ ïï
ï è Heat ø ï ï è 0.65 ø
E Heat = í ý ´ E F1 = í ý ´ 3356 kgCO2 eq = 1678kgCO2 eq
ï æç P1 ö÷ æç Heat öï
÷ ï æ 5MW ö æ 10.83MW ö ï
ç ÷+ç ÷
ïç e ÷ + ç e ÷ï ïî è 0.3 ø è 0.65 ø ïþ
î è P1 ø è Heat øþ

Development of efficiency factors for H1 and P2 is complicated by the fact that the CHP
system incorporates two fuel inputs (F1 and F2). Steam energy produced in the HRSG is
derived from a combination of waste heat from the gas-fired turbine (heat that originated
from part of the energy in fuel stream F1) and supplemental firing of natural gas (often
termed a duct burner). In allocating emissions associated with operating the HRSG, the
exhaust from the gas turbine is treated as a fuel and the emissions allocated to this stream
(EHeat) are added to the emissions associated with F2 (EF2), and these total emissions (EF2) are
allocated between H1 and P2.

There are differing efficiencies associated with converting each of these two energy sources
into steam in the HRSG. The mill has information that indicates that the efficiency of the
HRSG in converting the heat in the turbine exhaust gas into steam energy is 80%. The
efficiency associated with combustion of the auxiliary fuel in the duct burner is 100% (this is
typically true of supplementally fired HRSGs). This information can be used to develop an
overall efficiency of the HRSG as follows:

æ 10.55 MW ö æ 10.83MW ö
eff B = 100% ´ ç ÷ + 80% ´ ç ÷ = 90%
è 10.55MW + 10.83MW ø è 10.55MW + 10.83MW ø

It is assumed that the efficiency associated with H1 is equivalent to that of producing steam in
the HRSG (Hs), 90%. The mill has information indicating that the efficiency of the back
pressure steam turbine in converting expansion into mechanical work (isotropic expansion
efficiency) is 75%, and the generator which converts the mechanical work into electrical
power is 95% efficient. Therefore, the efficiency of producing electrical power output P2 is:

(effB) ´ (effturbine) ´ (effgenerator) = (0.9) ´ (0.75) ´ (0.95) = 0.64, or 64%.

E F 2' = E F 2 + E Heat = 2126 + 1678 = 3804 kgCO2 eq

ì æ 3MW ö ü
ïï ç ÷ ïï
E P2 = í
è 0.64 ø
ý ´ 3804 kgCO2 eq = 835kgCO 2 eq
ï æç 3MW ö÷ + æç 15MW öï
÷
ïî è 0.64 ø è 0.9 ø ïþ

ì æ 15MW ö ü
ïï ç ÷ ïï
E H1 = í
è 0.9 ø
ý ´ 3804 kgCO2 eq = 2969kgCO 2 eq
ï æç 3MW ö÷ + æç 15MW öï
÷
ïî è 0.64 ø è 0.9 ø ïþ

E-7
The following table presents a summary of emissions and emission factors for the three
outputs in this CHP example.

Total Efficiency Steam Steam CO2 CO2


Energy Temp. Press. Emissions Emission Factor
(MW) (°C) (bar) (kg CO2) (kg CO2/MWh)
P1 (Electricity) 5 0.3 N/A N/A 1678 336
P2 (Electricity) 3 0.64 N/A N/A 835 278
H1 (Steam) 15 0.9 170 7 2969 198

Total 5482

2.2. Heat Content Method

In the heat content method all of the energy in electrical power is considered useful;
however, only the fraction of the total energy in steam (or hot water) that can be used for
process heating is considered useful. Furthermore, it is assumed that the steam is used for
indirect heating, with condensates returned to the CHP system. Alternatively, if the
condensates are not returned or if a hot water output stream is considered in the allocation,
reference conditions other than those shown below may be used (e.g., the temperature and
pressure of boiler feed water). Therefore, the useful energy content of steam can be
calculated using Equation 5:

Useful Energy = Fi ´ (H i - H ) (Eq. 5)


ref

Where: Fi = the mass of steam in tonnes (1000 kg)


Hi = the specific enthalpy of steam flow i, in kJ/kg
Href = the specific enthalpy at reference conditions (corresponding to returned
condensates, assume at 100°C and 1 atm pressure)

If the quantity of steam (or hot water) is given in terms of total energy, the corresponding
mass of steam (or hot water) can be calculated using Equation 6:

total energy
Fi = (Eq. 6)
Hi

As an example, consider a CHP system which emits a total of 174,000 tonnes of carbon
dioxide per year with total energy outputs as shown in Table 1. The useful energy content of
electricity is equivalent to the total energy, and for the three steam streams the useful energy
is calculated from Equation 5. Allocated carbon dioxide emissions and an emission factor
(ton CO2 per GJ of total energy) for each energy output are also shown in Table 1.

E-8
Table 1. Allocation of GHG Emissions Based on the Heat Content Method
Combined Heat and Power system total CO2 emissions = H = 174,000 tonnes
A B C D E F
Total Steam Steam Useful CO2 CO2
Energy Temp. Press. Energy Emissions Emission Factor
Eq. 1 E=H´Di/åD F=E/A
(GJ) (°C) (barg) (GJ) (tonne CO2) (tonne CO2/GJ)
Electricity 245 N/A N/A 245 14,167 57.8
Steam 1 1355 400 40 1178 68,120 50.3
Steam 2 1100 300 20 947 54,762 49.8
Steam 3 750 200 10 639 36,951 49.3
Total 3450 3009 174,000

2.3 Work Potential Method

In other applications, the steam generated in the CHP system may be used to drive
mechanical equipment. In these cases, the work potential method of allocating emissions
may be more appropriate. The work potential method is not appropriate for CHP systems
which include a hot water output stream. As in the heat content method, the work potential
method considers all the energy contained in electrical power to be useful, and a fraction of
the energy in steam to be useful. However, in the work potential method the useful energy
fraction of the total energy in steam corresponds to the maximum amount of work that could
be done by the steam in an open (flow), steady state, thermodynamically reversible process.
The thermodynamic term for this amount of work is the “availability” or the “exergy.” The
exergy of a particular stream (the useful energy parameter corresponding to the work
potential method) relative to a reference case can be computed using Equation 7:

Useful Energy = Fi ´ ìíé H i - (T + 273) ´ S i ù - é H - (T + 273) ´ S ù ü (Eq. 7)


î ê
ë ref ú
û ê
ë ref ref ref úû ýþ
where: Fi = the mass of steam in tonnes (1000 kg)
Hi = the specific enthalpy of steam flow i, in kJ/kg
Href = the specific enthalpy at reference conditions (corresponding to returned
condensates, assume at 100°C and 1 atm pressure)
Si = the specific entropy of steam flow i, in kJ/kg×K
Sref = the specific entropy at reference conditions
Tref = the temperature at reference conditions

Table 2 presents the allocated carbon dioxide emissions and emission factors for each of the
energy outputs for the same example CHP system computed by the work potential method.

E-9
Table 2. Allocation of GHG Emissions Based on the Work Potential Method
Combined Heat and Power system total CO2 emissions = H = 174,000 tonnes
A B C D E F
Total Steam Steam Useful CO2 CO2
Energy Temp. Press. Energy Emissions Emission Factor
Eq. 3 E=H´Di/åD F=E/A
(GJ) (°C) (barg) (GJ) (tonne CO2) (tonne CO2/GJ)
Electricity 245 N/A N/A 245 48,200 196.8
Steam 1 1355 400 40 320 63,000 46.5
Steam 2 1100 300 20 210 41,200 37.5
Steam 3 750 200 10 109 21,500 28.7
Total 3450 884 174,000

E - 10
ANNEX F

GREENHOUSE GASES FROM VEHICULAR TRAFFIC AND MACHINERY -


OVERVIEW OF METHODS IN EXISTING PROTOCOLS

1.0 OVERVIEW

National inventories of GHG emissions from mobile sources focus on highway travel, and
rail, air, and water transport, with highway travel being the most significant component of
mobile source emissions, by far. Some of the mobile emissions of interest to pulp and paper
mills, for example emissions from vehicles used at industrial facilities or in forestry
operations, get little or no attention. Some of the inventory documents suggest emission
factors for non-highway utility and construction vehicles that appear to be applicable to some
of these sources.

The WRI/WBCSD GHG Protocol distinguishes between direct and indirect emissions from
mobile sources based on the ownership or control of the vehicle. The WRI/WBCSD
Protocol Scope 1 reporting requirements include all direct emissions, regardless of where
they occur (WRI 2001a). Because corporate inventories often include both on-site and off-
site vehicular emissions, references are given in this annex that can be used to estimate
emissions from a variety of off-road vehicles and equipment that is sometimes used by the
forest products industry.

Companies interested in estimating the emissions from company-owned on-road vehicles can
use information from a variety of organizations, including IPCC and WRI/WBCSD (IPCC
1997c and WRI 2001d). The WRI/WBCSD calculation tools for transportation emissions are
available on the internet (WRI 2001d) and are briefly summarized here as well. The Excelâ
workbook that accompanies this report incorporates the transportation calculation tools from
WRI/WBCSD.

1.1 Carbon Dioxide

Essentially all protocols suggest that CO2 emissions from transportation vehicles and
equipment be calculated from fuel consumption and carbon content data. This is affirmed in
the Revised 1996 IPCC Guidelines (IPCC 1997c), the May 2000 IPCC Good Practices
document (IPCC 2000), the EMEP/CORINAIR Atmospheric Emission Inventory
Guidebook, second edition (EEA 1999), and the WRI/WBCSD GHG Protocol calculation
tools (WRI 2001d). A number of the protocols also give emission factors (kg CO2/vehicle
km) as a method for cross checking the estimates.

It is reasonable to expect that companies will be able to estimate the consumption and carbon
content of fuels used in on-site vehicles. Lacking site-specific information on the carbon
content of fuel, companies can use the values published by national authorities. For a
number of countries and fuels, these values are listed in Annex A of these tools.

F-1
In some cases, a authorities issue a single emission factor shown in CO2-equivalents that
incorporates emissions of CH4 and N2O. In the Australian Greenhouse Challenge, for
instance, the emission factors not only include all three GHGs, they also include the impact
of upstream emissions from fuel extraction, processing, and transportation (i.e., “full fuel
cycle emissions”).

1.2 Methane and Nitrous Oxide

1.2.1 IPCC’s Revised 1996 Guidelines and May 2000 Best Practices Document

The Reference Manual of IPCC’s Revised 1996 Guidelines contains emission factors for
“surface non-road sources.” (IPCC 1997c, page 1.88) The Revised 1996 Guidelines
include emission factors published by EMEP/CORINAIR Atmospheric Emissions Inventory
Handbook, most recently updated in 1996, and by the US EPA. Both sets of emission factors
are shown in the following tables.

Table 1. Fuel Consumption-Based N2O and CH4 Emission Factors


for Non-Road Mobile Sources and Machinery (IPCC 1997c)
(IPCC Revised 1996 Guidelines taken from EMEP/CORINAIR)
Source and Engine Type g N2O/kg fuel g N2O/MJ g CH4/kg fuel g CH4/MJ
Forestry – diesel 1.3 0.03 0.17 0.004
Industry – diesel 1.3 0.03 0.17 0.004
Railways – diesel 1.2 0.03 018 0.004
Industry – gasoline 4 stroke 0.08 0.002 2.2 0.05
Forestry – gasoline 2 stroke 0.02 0.0004 0.04 7.7
Industry – gasoline 2 stroke 0.02 0.0004 0.05 6.0

1.2.2 EMEP/CORINAIR Atmospheric Emissions Inventory Guidebook (EEA 1999)

The EMEP/CORINAIR Atmospheric Emissions Inventory Guidebook contains a second set


of emission factors that are based on the power output of the engine. These emission factors
are presented in a way that allows them to be adjusted based on the engine design and the age
of the engine. They can be used to estimate emissions from all fossil fuel fired engines. The
emission factors and the needed adjustment factors are shown below (EEA 1999).

F-2
Table 2. N2O and CH4 Emission Factors for Non-Highway Vehicles (IPCC 1997c)
(IPCC Revised 1996 Guidelines taken from US EPA)

Source and Engine Type g N2O/kg fuel g N2O/MJ g CH4/kg fuel g CH4/MJ
Ships and boats
Residual 0.08 0.002 0.23 0.005
Distillate 0.08 0.002 0.23 0.005
Gasoline 0.08 0.002 0.23 0.005
Locomotives
Residual 0.08 0.002 0.25 0.006
Diesel 0.08 0.002 0.25 0.006
Coal 0.08 0.002 0.25 0.006
Farm Equipment
Gas/Tractor 0.08 0.002 0.45 0.011
Other Gas 0.08 0.002 0.45 0.011
Diesel/Tractor 0.08 0.002 0.45 0.011
Other Diesel 0.08 0.002 0.45 0.011
Construction
Gas Construction 0.08 0.002 0.18 0.004
Diesel Construction 0.08 0.002 0.18 0.004
Other Non-highway
Gas Snowmobile 0.08 0.002 0.18 0.004
Gas Small Utility 0.08 0.002 0.18 0.004
Gas Heavy Duty Utility 0.08 0.002 0.18 0.004
Diesel Heavy Duty Utility 0.08 0.002 0.18 0.004

F-3
Table 3. CORINAIR Engine Output-Based N2O and CH4 Emission Factors
for Non-Road Mobile Sources and Machinery (EEA 1999)
. .
N2O CH4
Source and Engine Type/Size (g/kWh) (g/kWh)
Baseline factors
Diesel engines 0.35 0.05
2-stroke gasoline 0-2 kW 0.01 6.60
2-stroke gasoline 2-5 kW 0.01 3.55
2-stroke gasoline 5-10 kW 0.01 2.70
2-stroke gasoline 10-18 kW 0.01 2.26
2-stroke gasoline 18-37 kW 0.01 2.01
2-stroke gasoline 37-75 kW 0.01 1.84
2-stroke gasoline 75-130 kW 0.01 1.76
2-stroke gasoline 130-300 kW 0.01 1.69
4-stroke gasoline 0-2 kW 0.03 5.30
4-stroke gasoline 2-5 kW 0.03 2.25
4-stroke gasoline 5-10 kW 0.03 1.40
4-stroke gasoline 10-18 kW 0.03 0.96
4-stroke gasoline 18-37 kW 0.03 0.71
4-stroke gasoline 37-75 kW 0.03 0.54
4-stroke gasoline 75-130 kW 0.03 0.46
4-stroke gasoline 130-300 kW 0.03 0.39
4-stroke LPG 0.05 1.0
Pollutant weighting factors for diesel engines (multiply baseline factors shown above
by these values)
Naturally aspirated direct injection 1.0 0.8
Turbo-charged direct injection 1.0 0.8
Intercooled turbo-charged direct injection 1.0 0.8
Intercooled turbo-charged prechamber injection 1.0 0.9
Naturally aspirated prechanger injection 1.0 1.0
Turbo-charged prechamber injection 1.0 0.95
Degradation factors (increase emission factors calculated above by these values)
Diesel engines 0% per year 1.5% per year
2-stroke gasoline engines 0% per year 1.4% per year
4-stroke gasoline and LPG engines 0% per year 1.4% per year

F-4
1.2.3 Australia – Greenhouse Challenge – Factor and Methods Workbook (AGO 2001b)

Manufacturers participating in the Australian Greenhouse Challenge estimate emissions


using emission factors that include CO2, CH4, and N2O, and also address full fuel cycle
emissions (i.e., they include upstream emissions from fuel extraction, processing and
transport) (AGO 2001b).

1.2.4 Finland - Greenhouse gas emissions and removals in Finland – April 9, 2001

The Finnish inventory document, “Greenhouse gas emissions and removals in Finland –
April 9, 2001,” contains a list of emission factors for “small scale combustion,” most of
which are from the CORINAIR Emission Inventory Handbook. Several of the factors of
interest to the forest products industry are listed below (Technical Research Center of
Finland 2001).

Table 4. Finnish Emission Factors for Forestry and Industrial Machinery


(Technical Research Center of Finland 2001)
.
Source Fuel CH4 N2O
(mg/MJ) (mg/MJ)
Forest machinery - diesel light fuel oil 3.9 31.5
Forest machinery – Otto engine gasoline 138.3 0.3
Construction machinery – diesel light fuel oil 4.1 30.8
Construction machinery – Otto engine gasoline 133.3 1.7
Other machinery – Otto engine gasoline 93.8 1.1
Other machinery – diesel light fuel oil 4.0 30.9
Other machinery – diesel LPG 63.9 3.2 .

1.2.5 United Kingdom

In the development of the UK national inventory, emission factors for off-road sources of
GHG emissions from fossil fuel combustion are mostly from the EMEP/CORINAIR 1996
Emission Inventory Handbook (AEA Technology 2001).

1.2.6 Canada - FPAC Greenhouse Gas Action Plan Guidelines – 2001

FPAC does not include transportation related emissions in its guidance to mills because
Environment Canada has notified FPAC that “companies should not calculate emissions for
transportation fuels as these emissions are already accounted for in Canada’s GHG
inventory…” (FPAC 2001). The Canadian government inventory document uses the
IPCC/EMEP/CORINAIR non-road vehicle emission factors from the 1996 Revised IPCC
Guidelines, after converting them to a fuel volume basis (i.e., grams of emission per litre of
fuel) (IPCC 1997c).

F-5
1.2.7 WRI/WBCSD Greenhouse Gas Protocol and supporting documents

Noting that CH4 and N2O emissions “comprise a relatively small proportion of overall
transportation emissions,” the WRI/WBCSD Protocol includes only CO2 emissions from
mobile sources. Companies are given the option of estimating CH4 and N2O emissions from
mobile sources (WRI 2001d). The following tables contain the WRI/WBCSD GHG
Protocol default emission factors according to fuel use and distance traveled (WRI 2001d).

Table 5. Default Emission factors and heating values for different transportation fuels
(WRI 2001d)
Fuel type Based on Lower Heat Value (derived) (derived) (derived)
kg CO2 / GJ GJ / litre GJ/ US GJ/Imp GJ / litre GJ/tonne
gallon Gal
Gasoline / petrol 69.25 0.0344 0.1302 0.1564 0.0344 43.5674
Kerosene 71.45 0.0357 0.1351 0.1623 0.0357 44.0768
Jet Fuel 70.72 (EIA) 44.5900
Aviation gasoline 69.11 (EIA) 0.0343 0.1299 0.1561 0.0343
Diesel 74.01 0.0371 0.1404 0.1687 0.0371 44.1667
Distillate fuel oil No.1 74.01 0.0371 0.1404 0.1687 0.0371 43.9211
Distillate fuel oil No.2 74.01 0.0371 0.1404 0.1687 0.0371 43.9211
Residual Fuel oil#4 74.01 0.0379 0.1435 0.1723 0.0379
Residual Fuel oil#5 77.30 0.0397 0.1503 0.1805 0.0397 39.9535
Residual Fuel oil#6 77.30 0.0405 0.1533 0.1841 0.0405 40.7586
LPG 63.20 0.0249 0.0942 0.1132 0.0249 45.9779
Lubricants 73.28 0.0382 0.0382
Anthracite 98.30 0.02860 GJ/kg 27'177 GJ / kg
Bituminous coal 94.53 0.0302 - 0.0326 GJ/kg 28'724 - 30'999 GJ / kg
Butane 0.0258 0.0258
Propane 62.99 (EIA) 0.0240 0.0240 47.3373
Anthracite 98.30 23.53 *
Bituminous coal 94.53 23.53 *
Sub-bituminous coal 96.00 23.53 *
Wood, wood waste 100.44 (EIA)
Natural gas 56.06 0.039 GJ / standard cubic meter 0.039 GJ / standard cubic
meter

F-6
Table 6. Default Fuel economy factors for different types of mobile sources and activity
data (WRI 2001d)
Vehicle Type Liters/100km mpg gram CO2 / km
New small gas/electric hybrid 4.2 56 100.1
Small gas auto, hghwy 7.3 32 175.1
Small gas auto, city 9.0 26 215.5
Medium gas auto, hghwy 7.8 30 186.8
Medium gas auto, city 10.7 22 254.7
Large gas automobile, hwy 9.4 25 224.1
Large gas automobile, city 13.1 18 311.3
Medium Station wagon, hwy 8.7 27 207.5
Med Station wagon, city 11.8 20 280.1
Mini Van, hwy 9.8 24 233.5
Mini Van, city 13.1 18 311.3
Large Van, hwy 13.1 18 311.3
Large Van, city 16.8 14 400.2
Mid size. Pick-up Trucks, hwy 10.7 22 254.7
Pick-up Trucks, city 13.8 17 329.6
Large Pick-up Trucks, hwy 13.1 18 311.3
Large Pick-up Trucks, city 15.7 15 373.5
LPG automobile 11.2 21 266
Diesel automobile 9.8 24 233
Gasoline light truck 16.8 14 400
Gasoline heavy truck 39.2 6 924
Diesel light truck 15.7 15 374
Diesel heavy truck 33.6 7 870
Light motorcycle 3.9 60 93
Diesel bus 35.1 6.7 1035

F-7
ANNEX G

GREENHOUSE GASES FROM WASTE MANAGEMENT AT PULP AND PAPER


MILLS – RECOMMENDED APPROACH AND REVIEW OF EXISTING METHODS

1.0 RECOMMENDED APPROACH FOR ESTIMATING CH4 EMISSIONS


FROM LANDFILLS – Note: Much of Sections 1.0 through 1.2.1 is identical to
Section 14 in the Calculation Tools Report

These calculation tools have been developed assuming that many companies will include
mill-owned landfills within the inventory boundaries. These tools can also be used in cases
where a mill’s process waste is being disposed in a municipal solid waste landfill and the
company is interested in estimating the mill’s contribution to the municipal landfill
emissions. The reporting format, however, has been prepared assuming that only the
emissions from company-owned landfills will be reported.

As is the case with most widely accepted protocols, only CH4 emissions are addressed in
these tools because the CO2 from landfills is composed of biomass carbon (not counted as a
greenhouse gas) and the N2O emissions are assumed to be negligible.

In Table 1 of the main body of this report, an emission factor for landfilled waste is
presented. This factor is based on a number of conservative assumptions and, in most cases,
is expected to be higher than the actual emissions attributable to landfilled mill waste. The
emission factor can be useful, however, in deciding whether landfill emissions are material to
the results of the inventory. For preparing an estimate to use in the inventory results,
however, these calculation tools recommend the methods, described below, all of which are
contained in the Excelâ workbook that accompanies this report.

1.1 Using Data from Landfill Gas Collection Systems

In some cases, company landfills are capped with low permeability cover material and the
landfill gas is collected. In many of these situations, the amounts of methane collected and
destroyed can be estimated from site-specific data. IPCC recommends that this information
be used by subtracting the amounts of methane destroyed from the amounts of methane that
the company estimates are generated by the landfill, using methods described below. The
problem with this approach is that, because of the large uncertainties in estimating methane
generation, the amounts burned (which were measured) could easily be greater than the
amounts generated (which were estimated), resulting in a negative release. It is equally
possible that the comparison of estimated generation rates to measured collection rates could
suggest impossibly low collection efficiencies, due only to the uncertainties in estimating
methane generation.

An alternative approach is available to companies that measure the amounts of methane


captured in the collection system. The alternative approach is to estimate the collection
efficiency of the collection system and then back-calculate the amounts of methane

G-1
generated. For instance, if a mill with a capped landfill has determined that its collection
system collects 90 tons of methane per year and the mill estimates that the collection
efficiency is 90%, it means that 100 tons of methane were generated.

The problem with this approach is that the effectiveness of landfill gas collection systems is
variable and uncertain. Reported collection efficiencies range from 60 to 85% (USEPA
1998). This variability and uncertainty has caused IPPC to take the position that “the use of
undocumented estimates of landfill gas recovery potential is not appropriate, as such
estimates tend to overestimate the amount of recovery” (IPCC 2000). Nonetheless, this
approach is built around a measured value – the amount of gas collected. For this reason, it
is reasonable to expect that in some cases it may yield more accurate estimates than IPCC’s
default methodology. This is especially true for mill landfills because of the limited data for
deriving the parameter values needed to use IPCC’s default methodology on pulp and paper
mill wastes.

Therefore, in these calculation tools, it is recommended that where mill landfills are covered
with low permeability caps and equipped with landfill gas collection systems constructed and
operated to normal standards, the methane generation rates should be back calculated from
measurements of the amounts of methane collected and estimates of collection efficiency. A
default collection efficiency of 75% has been used by some authorities and is recommended
here, unless site-specific collection efficiency data are available (USEPA 1998).

These calculation tools also assume that all of the methane that is captured and burned is
converted to biomass CO2 and therefore does not have to be included in the inventory.

Using these default values and assumptions, estimates of methane generation can then be
developed by using the following equation.

CH4 (m3/y) released to the atmosphere = [(REC / FRCOLL) * (1 – FRCOLL) * FRMETH *


( 1 – OX )] + [ REC * FRMETH * ( 1 – FRBURN)] (Eq.1)

where:
REC = amount of landfill gas collected, determined on a site specific basis, m3/y
FRCOLL = fraction of generated landfill gas that is collected, default is 0.75
FRMETH = fraction methane in landfill gas, default is 0.5
OX = fraction of methane oxidized in the surface layer of the landfill, default is 0.1
FRBURN = fraction of collected methane that is burned, site specific determination

1.2 Estimating Landfill Methane Emissions at Mill Landfills without Gas Collection
Data

1.2.1 The Simplified First Order Decay Approach

Where the approach described above cannot be used, it is recommended that companies
employ the first order decay model approach for estimating landfill gas emissions using
parameter values derived for pulp and paper mill landfills. This approach is the default

G-2
method recommended by IPCC and is used by a number of national authorities (IPCC 2000).
It can be used to estimate CH4 emissions from active and inactive landfills and involves the
following analysis.

In cases where the annual deposits are constant (or assumed to be) IPCC’s default method
reduces to two equations. This approach should be adequate unless the amounts or types of
waste being landfilled have changed significantly from year to year (e.g., a new deinking mill
is built on-site), or the landfill design or operation have been changed in a way that would
significantly impact methane generation or release (e.g., a gas collection system is installed).
The simplified approach is as follows.

CH4 (m3/y)generated from all waste in the landfill = R L0 (e-kC - e-kT) (Eq.2)

where:
R = average amount of waste sent to landfill per year, Mg/y
L0 = ultimate methane generation potential, m3/Mg waste
k = methane generation rate constant, 1/y
C = time since landfill stopped receiving waste, y
T = years since landfill opened, y
(Note: R and L0 can be in units of wet weight, dry weight, degradable organic
carbon, or other units but the units for R and L0 must be the same.)

Not all methane that is generated is subsequently released to the atmosphere. To estimate
atmospheric releases, use the result from equation 2 in equation 3 below. For landfills with
modern gas collection and combustion systems but no measurements of quantities of gas
collected, the amount of methane recovered can be assumed to be 75% of that generated
(USEPA 1998).

CH4 (m3/y) released =


[(CH4 generated–CH4 recovered)*(1–OX)]+[CH4 recovered*(1-FRBURN)] (Eq.3)
where:
CH4 generated = from Equation 2
CH4 recovered = amount of methane collected, site specific determination
OX = fraction oxidized in the surface layer of the landfill before escaping, usually
assumed to be 0.1
FRBURN = fraction of collected methane that is burned, site specific determination

If the amounts being landfilled have changed significantly or if the landfill design has been
altered so that some of the parameter values would have changed substantially, a more
involved approach may be needed. To deal with these more complicated situations, many
protocols recommend modeling the gas generated annually from each year’s deposits and
them summing the amounts that are predicted to occur in the current year. This more
detailed analysis is described below.

G-3
A number of sources for the parameter values L0 and k needed in these equations are shown
below. Unfortunately, the values vary considerably from one protocol to the next and the
values are based on very few data.

1.2.2 Detailed First Order Decay Approach

To allow year-to-year variations in the amounts of waste sent to landfill, IPCC suggests a
variation of this approach shown below. Using this variation, starting in year one, you
calculate how much methane will be generated in each subsequent year by waste deposited in
that year using the following equations.

CH4 generated in a given year by waste


deposited in an earlier year (m3/y) = k Ry L0 (e-k[T-Y]) (Eq.4)

Where
k = methane generation rate constant, 1/yr
RY,= the amount of waste sent to landfill in year Y, Mg/yr
L0,= ultimate methane generation potential, m3/Mg waste
T = year for which emissions are being estimated given in terms of
years since the landfill opened
Y = year after landfill opened that waste was disposed
So (T-Y) is equal to the number of years the waste has been in place
prior to the year for which emissions are being estimated.

CH4 (m3/y) released =


[(CH4 generated–CH4 recovered)*(1–OX)]+[CH4 recovered*(1-FRBURN)] (Eq.5)
where:
CH4 generated = from Equation 5
CH4 recovered = amount of methane collected, site specific determination
OX = fraction oxidized in the surface layer of the landfill before escaping, usually
assumed to be 0.1
FRBURN = fraction of collected methane that is burned, site specific determination

To perform the calculations you need to estimate how much waste was deposited every year
since the landfill was opened. IPCC indicates that for very old landfills, it is possible to limit
the retrospective period to one starting at least three waste degradation half-lives before the
current year. Given the slow degradation observed in many mill sludges, 25 years is
probably the minimum that would satisfy this criterion. For each year’s deposit, you then
estimate for that year and each following year the amount of methane released. In
subsequent years, the amount of methane released is the sum of the amounts estimated from
each prior year’s deposits that were projected to occur in that year.

The calculations work like this. In year 1 you deposit amount A and estimate that in years 1,
2, 3, … it will release X1, X2, X3, … tons of methane, respectively. The reported emissions
for year one are X1 tons of methane. In year 2 you deposit amount B and estimate that in

G-4
years 2, 3, 4, … it will release Y2, Y3, Y4, …tons of methane, respectively. The emissions
reported for year 2 are X2 plus Y2 tons methane. In year 3, you deposit amount C and
estimate that in years 3, 4, 5, … it will release Z3, Z4, Z5, … tons of methane, respectively.
The reported emissions for year 3 are X3 plus Y3 plus Z3 tons of methane. This process
repeats itself every year.

The values for k and L0 are the same as those used in the simplified first order approach.

1.2.3 First order methane generation rate constant, k

Even for municipal waste landfills, there is large uncertainty about the proper first order rate
constant. The correct value for pulp and paper mill landfills is even more uncertain. The
guidance offered by the sources reviewed in this study is summarized as follows.

IPCC – k varies from 0.005 to 0.4 per year, with a default of 0.05/yr for MSW (IPCC
1997c)

UK – k varies from 0.05/yr for slowly degrading waste to 0.185/yr for rapidly
degrading waste (AEA Technology 2001)

Sweden – k equals 0.092/yr for all landfills (Swedish Environmental Protection


Agency 2001)

Canada – k for wood waste landfills equals 0.01/yr, varies from 0.003/yr to 0.028/yr
for MSW landfills (Environment Canada 1999)

EPA – k equals 0.04/yr for areas receiving at least 25 inches (63.5 cm) of
precipitation and 0.02/yr for drier areas (for MSW landfills) (USEPA 1998)

1.2.4 Ultimate methane generation potential, L0

Again, there is a great deal of variability in the values being used for L0. The following
parameter values are for MSW unless otherwise indicated. It is also important to note that L0
can be expressed as wet weight, dry weight, and a number of other ways. Any form is
acceptable, but the units of L0 and R (the amount of waste disposed) must be the same.
Values given for municipal solid waste are often for wet waste as disposed.

IPCC – The sources cited by IPCC indicate that L0 can vary from less than 100 to
over 200 m3/Mg. The following equation is provided for calculating a site-
specific L0 (IPCC 1997c).

L0 = (DOC, fraction degradable organic carbon in waste) x


(DOCf , fraction of DOC that degrades into landfill gas) x
(16/12, to convert carbon to methane) x
(F, fraction CH4 in gas from a managed landfill, default
value is 0.5) x (MCF, amount of methane in landfill gas

G-5
relative to a managed landfill)

For MSW landfill default values, IPCC recommends the following (IPCC 1997c,
2000)
DOC – the default values for different countries range from 0.08 to
0.21 but site specific determinations are recommended
DOCf – the default range, if the DOC includes lignin, is 0.5 to 0.6
F – default is 0.5
MCF - 1.0 for modern managed landfills, 0.4 for shallow unmanaged
landfills (less than 5 meters deep), and 0.8 for deeper
unmanaged landfills is 0.8.

UK – The equation above is used to calculate L0. The DOC for different types of
waste was determined form a national study. DOCf was assumed to be 0.6. F
was usually 0.5 but 0.3 was used for old shallow sites. MCF was assumed to be
1.0 (AEA Technology 2001).

Sweden – Sweden uses an L0 of 45 kg CH4/ton of waste specifically for pulp and


paper mill sludge landfills. This is equivalent to 63 m3/Mg (Swedish
Environmental Protection Agency 2001).

Finland – Although Finland does not use the first order model approach for landfill
methane, the approach it uses still requires it to estimate L0. Finland uses the
IPCC equation shown above for L0. It uses the following values for the
variables in the equation (Technical Research Center of Finland 2001).

DOC = 0.4 for paper and cardboard, wet weight basis


= 0.3 for wood and bark, wet weight basis
= 0.1 for deinking waste, wet waste basis (definition is uncertain since
deinking sludge is listed separately)
= 0.45 forest industry sludge-unspecified, dry weight basis (assumed 30%
solids)
= 0.3 deinking sludge, dry weight basis (assumed 30% solids)
= 0.3 forest industry fiber sludge, dry weight basis (assumed 30% solids)

DOCf = Finland uses 0.5 to reflect the low temperature and less-than-optimal
conditions for decomposition in Finnish landfills

MCF = 0.7 (Finland assumes that one-half of waste goes to small landfills
with MCF=0.4 and with the rest going to large landfills with
MCF=1)

F = 0.5

G-6
Putting these together and using a range in DOC of 0.3 to 0.45 for mill wastes
on a dry weight basis and assuming an MCF of 1, the calculated range for L0 is
0.1 to 0.15 kg CH4/kg dry waste or 140 to 210 m3/Mg .

Canada – To calculate the methane potential for wood waste landfills, Canada uses an
L0 of 118 kg CH4/ton wood waste which converts to 165 m3/Mg. The value
for MSW landfills is 117 kg CH4/ton (Environment Canada 1999).

US EPA – EPA’s compilation of emission factors (AP-42) indicates that a value of


100 m3/Mg is recommended as the default factor for most MSW landfills
(USEPA 1998).

1.2.5 Recommended Default Values for k and L0

For situations where pulp and paper mill wastewater treatment sludge is a major constituent
of the waste, reasonable values for the rate constant, k, fall in the range of 0.01/yr to 0.1/yr
while those for L0 fall between 50 m3/Mg and 200 m3/Mg. NCASI is currently conducting
research that should narrow these ranges. Initial indications are that the amounts of gas
generated in mill landfills are less than would be predicted using parameter values developed
for municipal solid waste (NCASI 1999). With this knowledge, it is recommended that until
the current research is completed, and unless companies have country-specific or site-specific
factors that are more appropriate for their mill wastes, companies use the parameters values
shown in Table 1.

Table 1. Recommended Default Values for k and L0 for Estimating Mill Landfill Methane
Emissions
Parameter Default Value
k 0.03 y-1
L0 100 m3/Mg dry weight of waste

2.0 RECOMMENDED APPROACH FOR ESTIMATING CH4 EMISSIONS


FROM THE ANAEROBIC TREATMENT OF WASTEWATER OR SLUDGE
– Note: Most of Sections 2.0 through 2.2 is identical to Section 15 in the
Calculation Tools Report.

Most existing GHG protocols address GHG emissions only from anaerobic treatment and
digestion processes. Therefore, these calculation tools have been developed assuming that
emissions from other types of wastewater and sludge treatment processes are negligible.
Although aerobic and facultative treatment systems may have zones with depleted dissolved
oxygen, methane generation rates in aerated stabilization basins, activated sludge systems,
and their associated retention ponds would be expected to be much less than in anaerobic
systems. In any event, due to lack of data, it is not possible to reasonably estimate emissions
from aerobic and facultative treatment operations.

G-7
Even for anaerobic systems, only CH4 emissions need to be estimated because (a) the CO2
emitted from wastewater and sludge treatment operations contains biomass carbon which is
not included in most GHG protocols and (b) other protocols assume that N2O emissions, if
any, occur after the wastewater is discharged.

2.1 Anaerobic Treatment Operations where Off-Gases are Captured

In many cases, anaerobic treatment systems are covered and the gases are collected and
burned. One of the purposes of these collection systems is the prevention of odors, and to
accomplish this objective, the systems must be highly efficient. For purposes of a GHG
inventory it is reasonable to assume, therefore, that where methane emissions from anaerobic
treatment operations are captured and burned, the collection and destruction is complete and
no methane is emitted. If circumstances at a mill suggest that non-trivial amounts of methane
are escaping collection, the mill may need to undertake efforts to account for these releases,
but such circumstances are expected to be unusual at mills that collect and burn these gases.

Of course, if the gases are collected but released to the atmosphere rather than being burned,
they should be included in the inventory.

2.2 Anaerobic Treatment Operations where Off-Gases are Released to the


Atmosphere

Where off-gases from anaerobic treatment operations are not collected and burned, it is
necessary to estimate the releases of methane to the atmosphere. In some cases, for instance
where the gases are released through a vent in a covered vessel, the releases can be measured
directly. In most other cases, they must be estimated.

These calculation tools suggest the use of the IPCC default methodology as described in the
May 2000 Good Practices document and shown in the equation below (IPCC 2000).
Although the IPCC document allows the following equation to be applied to systems that are
not completely anaerobic (by multiplying the result of the following equation by an arbitrary
adjustment factor of less than one) there are no data currently available to support the
selection of the adjustment factor. It is recommended, therefore, that methane emissions only
be estimated from anaerobic treatment or sludge digestion systems until such time as factors
for other types of systems are available.

Anaerobic Treatment Plant Methane Emissions (kg/y) = ( OC x EF ) – B (Eq.6)

where:
OC = BOD or COD of the feed to the anaerobic system, kg/year
EF = emission factor, default values = 0.25 kg CH4/kg COD in the
feed or 0.6 kg CH4/kg BOD in the feed (or another
BOD-based factor developed by multiplying the COD-
based factor of 0.25 kg CH4/kg COD by the site-specific
COD/BOD ratio)

G-8
B = methane captured and burned, kg CH4/year, determined on a site
specific basis

If the solids are handled separately, emissions from sludge digestion would be calculated
using equation 7 below. In cases where sludge is burned, it is included in the calculations for
GHG emissions from biomass burning, discussed elsewhere.

Anaerobic Sludge Digestion Plant Methane Emissions (kg/y) = (OCs x EFs) - B (Eq.7)

where:
OCs = organic content of the sludge
EFs = emission factor, in units consistent with OCs - IPCC’s default
value is 0.25 kg CH4/kg COD in the sludge feed
B = methane captured and burned, kg CH4/yr, determined on a site-
specific basis.

Under most protocols, emissions of N2O from wastewater are normally assumed to take place
after wastewater is discharged into receiving waters. These emissions, therefore, are not
addressed in these tools.

3.0 OVERVIEW OF EXISTING APPROACHES

3.1 IPCC – Revised 1996 IPCC Guidelines for National Greenhouse Gas Inventories
(IPCC 1997c) and Good Practice Guidance and Uncertainty Management in
National Greenhouse Gas Inventories (IPCC 2000)

Note: In the following discussion, some of the symbols used are different than those used in
the IPCC documents. This has been done in an attempt to eliminate potential confusion
among variables having similar symbols in the IPCC documents.

3.1.1 Landfills – The All-in-One-Year Approach

The IPCC default methodology for estimating methane emissions from landfills is limited to
municipal solid waste landfills. Chapter 6 of the 1996 IPCC Guidelines Reference Manual
(IPCC 1997c) and Chapter 5 of the May 2000 Good Practices document (IPCC 2000) outline
two general approaches for estimating landfill emissions. The first approach assumes that all
organic matter degrades in the year it is placed in the landfill while the second approach uses
a first order model to estimate the amounts released over time. The Good Practices
document indicates that the first order decay approach should be used where possible (IPCC
2000).

The all-in-one-year approach starts with an estimate of the “degradable organic carbon”
(DOC) content of the waste going to landfill. IPCC provides the following default values for
certain large volume materials.

G-9
Table 2. Default DOC Values for Major Waste Streams from IPCC’s 1996 Revised
Methodology
(values are for “wet or fresh” materials in municipal solid waste)

Percent Degradable Organic Carbon,


Waste Stream % by weight
Paper and textiles 40
Garden and park wastes and other 17
non-food organic putrescibles
Food waste 15
Wood and straw waste (excluding 30
lignin)

Of the amount of carbon that is degradable organic carbon, only a fraction is converted into
landfill gas. This fraction is given the symbol DOCf. IPCC’s 1996 Guidance relies on a
simple model to generate a default value of 0.77 for DOCf but IPCC’s May 2000 “Good
Practice Guidance and Uncertainty Management” document indicates that this appears to be
too high unless the lignin carbon is excluded from the DOC value. The May 2000 document
goes on to say that a “good practice” default value of 0.5 to 0.6 should be used for DOCf in
cases where the lignin is included in the DOC unless better site-specific data are available
(IPCC 2000).

IPCC then applies a factor, the methane correction factor or MCF, which is intended to
account for the fact that landfill design and operation can influence the tendency of
degradable carbon to decompose into carbon dioxide rather than methane. The MCF is
simply a measure of the relative methane generation potential of unmanaged landfills
compared to managed landfills. A landfill is a “managed landfill” if it involves controlled
placement of waste, a degree of control over access and at least one of the following - cover
material, mechanical compacting, or leveling of waste. Managed landfills are used as the
baseline condition so the MCF is 1.0 for such landfills. The MCF for shallow unmanaged
landfills (less than 5 meters deep) is 0.4 while the factor for deeper unmanaged landfills is
0.8. At a managed landfill, the default assumption is that methane comprises 50% of the
landfill gas. The MCFs modify this assumption for unmanaged landfills; i.e., methane
comprises 40% of the gas from deep unmanaged landfills (0.8 x 50%) and 20% of the gas
from shallow unmanaged landfills (0.4 x 50%).

At many landfills, gas is trapped and burned, converting the carbon to biomass-CO2.
Methane generated within a landfill may also be converted to biomass-CO2 as it migrates
through the surface of the landfill. Because the CO2 formed from landfill methane is biomass
carbon, it is not included in IPCC emission inventories. In IPCC’s words, “[d]ecomposition
of organic material derived from biomass sources (e.g., crops, forests) which are regrown on
an annual basis is the primary source of CO2 released from waste. Hence, these CO2
emissions are not treated as net emissions from waste in the IPCC Methodology. If biomass
raw materials are not being sustainably produced, the net CO2 release should be calculated

G - 10
and reported under the Agriculture and Land-Use Change and Forestry Chapters.” (IPCC
1997c)

Overall, therefore, the IPCC all-in-one year approach involves the following calculations-

Methane generated =
(amount of waste sent to landfill) x DOC x DOCf x 16/12 x 0.5 x MCF (Eq.8)

Where
DOC = fraction degradable organic carbon in waste (based on the
same units as used to measure the amount of waste sent to
landfill)
DOCf = fraction of DOC that degrades into landfill gas
16/12 = conversion factor from carbon to methane
0.5 = fraction methane in gas from a managed landfill, default value
MCF = amount of methane in landfill gas relative to a managed
landfill (managed landfill MCF = 1)

Methane released = ( Methane generated – REC ) x ( 1 – OX ) (Eq.9)

Where:REC = amount of methane converted to CO2 by burning


OX = fraction of methane converted to CO2 by oxidation in the landfill
cover (default value is zero but IPCC’s May 2000 Good
Practices document indicates that a value of 0.1 can be used
for well managed landfills in industrialized countries)

3.1.2 Landfills – The First Order Decay Approach

IPCC suggests two approaches for modeling methane releases from landfills over time. The
first uses the average waste acceptance rate over the life of the landfill (see Section 1.2.1 of
this annex for a description) while the second considers each year’s waste separately (see
section 1.2.2 of this annex for a description) (IPCC 1997c).

3.1.3 Wastewater treatment and anaerobic sludge digestion

In aerobic treatment plants, most of the organic matter in mill wastewaters is either converted
to biological sludge or carbon dioxide. Because the carbon originated in biomass, the CO2
emitted from wastewater treatment is not included in GHG inventories. Methane and nitrous
oxide, however, can also be released during wastewater treatment. Methane, in particular, is
an important emission from anaerobic wastewater treatment and sludge digestion processes.
Methane and nitrous oxide are usually included in GHG inventories.

The Waste section of the 1996 Revised Guidelines contains a method for estimating N2O
emissions from “human sewage” (IPCC 1997c, page 6.28). The discussion in the Waste
section directs the reader to the Agriculture section of the Manual for more information. In
that section of the manual, IPCC indicates that three studies have examined nitrous oxide

G - 11
emissions from operating wastewater treatment facilities (IPCC 1997c, page 4.110). All of
the studies reported low N2O emissions. For this reason, in IPCC’s methodology, “N2O
associated with sewage treatment and land disposal is assumed to be negligible” and it is
further assumed that “all sewage nitrogen enters rivers and/or estuaries” where a portion of it
is converted into N2O (IPCC 1997c). In summary, IPCC’s guidance contains methods for
estimating the N2O released from human sewage once it is discharged, but assumes that N2O
emissions from the treatment plants are negligible. There is no discussion of N2O emissions
from pulp and paper or other industrial wastewater treatment plants.

Methane from wastewater treatment, however, receives much greater attention in IPCC’s
guidance documents. A diagram in the IPCC Good Practice document shows which types of
treatment processes have “the potential for CH4 emissions” (IPCC 2000, page 5.17). The
figure indicates that all aerobic treatment processes are outside of the scope of IPCC’s
guidelines due to the low potential for methane generation. The guidelines focus on
anaerobic wastewater treatment and anaerobic sludge digestion.

The Reference Manual of IPPC’s 1996 Revised Guidelines contains the following discussion
of pulp and paper industry waste treatment operations. (IPCC 1997c, page 6.16)

“Assessment of CH4 production potential from industrial wastewater streams is based


on the concentration of degradable organic matter in the wastewater, the volume of
wastewater, and the propensity of the industry to treat their wastewater in anaerobic
lagoons. “ [emphasis added]

Using these criteria, IPCC cites work by Doorn and Eklund suggesting that the paper and
pulp manufacturers are among the most likely to generate methane in wastewater treatment
(IPCC 1997c).

“Both the paper and pulp industry and the meat and poultry processing industries
produce large volumes of wastewater that contain high levels of degradable organics.
Additionally, both industries utilize large facilities that often have their own
wastewater handling systems. The meat and poultry processing facilities commonly
employ anaerobic lagoons to treat their wastewater, while the paper and pulp industry
is known to use lagoons.”

IPCC’s approach for estimating methane from wastewater treatment or sludge digestion is
similar to the all-in-one-year method used for estimating methane emissions from landfills.
Emissions are calculated using the following equation:

Methane emissions = (TOW or TOS) x B0 x MCF (eq.10)

Where:
TOW or TOS = measure of organic content of anaerobically treated
wastewater or sludge
B0 = CH4 per unit of organic matter, in units consistent with TOW or TOS

G - 12
MCF = fraction of methane not recovered or flared, varies from 0 to 1
depending on the treatment unit

First the amount of degradable substrate being fed to the treatment unit is determined. The
total organic (chemical oxygen demand, COD) in wastewater is given the symbol TOW
while the total organic in sludge is given the symbol TOS. The IPCC reference manual
contains some limited data that can be used to estimate pulp and paper mill wastewater COD,
but the data are not included in this report because the data available to individual companies
are expected to be of far greater quality (IPCC 1997c).

After dividing the untreated wastewater COD into TOW and TOS, the two streams are kept
separate in subsequent calculations. Landfill calculations for sludge are shown above, while
emissions from sludge digestion would be calculated using equation 10 above. In cases
where sludge is burned, it is included in the calculations for GHG emissions from biomass
burning, discussed elsewhere.

For industrial wastewaters, the maximum methane producing capacity is given the symbol B0
and is expressed in kg CH4/kg COD. IPCC’s 1996 Revised Guidelines give a default value
for B0 of 0.25 kg CH4/kg COD for wastewater and sludge. A footnote in the IPCC guidelines
explains that since the degradable organic matter in COD is the same material that is
measured as degradable BOD, the factor B0 will be 0.25 kg CH4/kg COD or BOD. This is
approximately true if the factors are based on the kg of BOD or COD removed, but the
Guidelines do not specify this, saying only that TOW and TOS are the total organic contents
in industrial wastewaters and sludges, respectively (IPCC 1997c).

The IPCC Good Practice document changes the 1996 Guidelines with the following
explanation. (IPCC 2000, page 5.17)

“Note that degradable carbon in organic waste can be measure in terms of either BOD
or COD. For typical domestic raw sewage, COD (mg/l) is 2 to 2.5 times higher than
BOD (mg/l). Therefore, it is important to use emission factors that are consistent
with the measure of degradable carbon being used. The IPCC Guidelines provide
only one default value of B0 that has to be applied to both COD and BOD. This is not
consistent with the observed differences between BOD and COD levels in raw
sewage. Given the differences in the amount of BOD and COD in wastewater this
can result in estimates of different emissions levels from the same amount of
wastewater depending on which measure is used. To ensure that the resulting
emission estimate from a given amount of wastewater is the same regardless of the
measure of organic carbon used, the COD-based value of B0 should be converted into
a BOD-based value via up-scaling with a default factor of 2.5. Thus, it is good
practice to use a default value of 0.25 kg CH4/kg COD or a default value of 0.6 kg
CH4/kg BOD.”

The important point is that the basis for the factors must match the measure of organic
content of the waste. In particular, one needs to know whether the factors are for BOD or
COD and whether they are based on the organic content of the untreated wastewater or the

G - 13
organic content removed during treatment. This can be especially important for many pulp
and paper mill wastewaters where the COD to BOD ratio can be very high.

A methane conversion factor (MCF) is again used to represent the methane generation
potential relative to a reference system. In this case, there are two reference systems. For
completely aerobic systems, the MCF is 0.0 while the MCF for completely anaerobic
systems is 1.0. Although the Reference Manual shows default MCFs for various countries,
IPCC suggests that experts be consulted in determining appropriate values for MCF (IPCC
1997c). In this review, no published values of MCF for aerobic or facultative treatment
systems were found.

3.2 Australia – Greenhouse Challenge Factor and Methods Workbook (AGO 2001b)

3.2.1 Landfills

The Greenhouse Challenge workbook uses the IPCC all-in-one-year approach for estimating
emissions from municipal solid waste landfills .

3.2.2 Industrial Wastewater Treatment

Total emissions from industrial wastewater are the sum of emissions from wastewater
treatment and sludge treatment and are calculated from the following equation, which is
essentially the same approach as suggested by IPCC for anaerobic wastewater and sludge
treatment operations (AGO 2001b). Although the AGO Greenhouse Challenge workbook
does not specifically indicate that the equation is to be used only for anaerobic treatment
operations, it is clearly not appropriate for aerobic or facultative treatment operations.

GHG emissions (t CO2-equiv.) =


[[((W x O x DOCw x (1-Fsl) x EFw)+(W x O x DOCsl x Fsl x EFsl))-R] x 21]/1000

= [W x O x [((DOCw x (1-Fsl) x EFw) +(DOCsl x Fsl x EFsl))-R] x 21]/1000 (Eq.11)

Where:Wastewater (W) = wastewater in m3 per ton of product


Output (O) = company output per annum in ton
DOCw = quantity in kilograms of COD/m3 wastewater
Fsl = fraction of organic degradable organic component removed as sludge,
default = 0.54
DOCsl = quantity in kilograms of COD/m3 sludge
EFw = methane emission factor for industrial wastewater, default = 0.25 kg
CH4/kg COD
EFsl = methane emission factor for industrial sludge, default = 0.25kg CH4/Kg
COD
R = recovered CH4 in the inventory year and measured in ton
21 = the global warming potential (GWP) of CH4

G - 14
3.3 Canada – FPAC’s 2001 Greenhouse Gas Action Plan Guidelines (FPAC 2001)
and Canada’s Greenhouse Gas Inventory, 1997 Emissions and Removals with Trends
(Environment Canada 1999)

The FPAC “2001 Greenhouse Gas Action Plan Guidelines” do not specifically include
emissions from waste management activities.

These emissions are included, however, in “Canada’s Greenhouse Gas Inventory, 1997
Emissions and Removals with Trends” (Environment Canada 1999).

Canada does not count CO2 produced by the decomposition of biomass carbon. Estimates
are made for CH4 and N2O emissions, however (Environment Canada 1999).

3.3.1 Landfills

Because the character of Canada’s landfills has been changing over time, Canada uses the
Scholl Canyon model (first order decay model) for estimating methane emissions from
landfills. This allows varying amounts of waste to be deposited every year and also allows
the decay rate to vary according to management practices and other factors. This is one of
the options described in the IPCC 1996 Revised Guidelines and the May 2000 IPCC Good
Practices document. The following description highlights only those aspects of the Canadian
approach that either involve Canada-specific parameter values or represent departures from
the IPCC approach (Environment Canada 1999).

Because of the large size of the forest products industry in Canada, the government
developed separate estimates of methane emissions from wood waste landfills. After
evaluating the recommendations developed by Canadian experts for “k” values to model
municipal solid waste (MSW) landfills, the government decided to use the lowest “k” value
for MSW in the major forest industry provinces to estimate emissions from wood waste
landfills. This “k” was 0.01/yr. The Canadian government considered the degradable carbon
content of wood waste and assumed that wood waste landfill gas would consist of 50%
methane to calculate a methane generation potential, L0, for wood waste of 118 kg CH4/tonne
of wood waste (Environment Canada 1999).

3.3.2 Wastewater treatment and sludge digestion

Canada only estimated GHG emissions from municipal wastewater treatment due to lack of
data on treatment of industrial wastewater (Environment Canada 1999).

Methane emissions from aerobic systems were assumed to be negligible, an assumption


consistent with IPCC’s guidance. Emissions from anaerobic systems were estimated using
an approach developed by Ortech International for Environment Canada in 1994. Using this
approach, it was estimated that 4.015 kg CH4/person/year could potentially be emitted from
wastewater treated anaerobically (Environment Canada 1999). This factor was multiplied by
the number of persons in each province and the fraction of the wastewater treated

G - 15
anaerobically in each province to estimate methane emissions from the anaerobic treatment
of municipal wastewater.

Canada used the IPCC default methodology for estimating N2O emissions from human
sewage. The IPCC methodology assumes that (a) negligible amounts of N2O are released
during treatment, and (b) all of the nitrogen in untreated human sewage is discharged to
rivers or estuaries, where a portion of the nitrogen is converted to N2O (Environment Canada
1999).

3.4 Finland - Greenhouse gas emissions and removals in Finland (Technical


Research Center of Finland 2001)

3.4.1 Landfills

The Finnish government’s estimates are based on IPCC’s all-in-one-year default method
(IPCC 1997c). The parameter values have been selected to represent conditions in Finland
(Technical Research Center of Finland 2001). The parameter values needed to estimate
emissions from pulp and paper mill landfills are shown below. See the description of the
IPCC all-in-one-year approach for more information on the variables and calculations.

Methane released per year =


([waste disposed per year x DOC x DOCf x MCF x F x 16/12] – R) - (1-OX) (Eq.12)

DOC = weight fraction Degradable Organic Carbon in waste (Varies by waste type
as shown below. Although not specified, the values for DOC strongly
suggest that lignin is included.)
= 0.4 for paper and cardboard, wet weight basis
= 0.3 for wood and bark, wet weight basis
= 0.1 for deinking waste, wet waste basis (definition is uncertain since
deinking sludge is listed separately)
= 0.45 forest industry sludge-unspecified, dry weight basis (assumed 30%
solids)
= 0.3 deinking sludge, dry weight basis (assumed 30% solids)
= 0.3 forest industry fiber sludge, dry weight basis (assumed 30% solids)

DOCf = 0.5 (This is the fraction of the degradable organic carbon that is degraded to
landfill gas. IPCC’s 1996 Revised Guidelines suggest a default of 0.77 but
the May 2000 Good Practices document revised this default value to 0.5 to 0.6
for DOCf in cases where the lignin is included in the DOC. Finland uses 0.5
to reflect the low temperature and less-than-optimal conditions for
decomposition in Finnish landfills.)

MCF = 0.7 (The Methane Correction Factor is essentially the landfill’s methane
generation potential relative to a “managed” landfill. Finland assumes that
one-half of waste goes to small landfills with MCF=0.4 and with the rest going
to large landfills with MCF=1)

G - 16
F = 0.5 (fraction of landfill gas that is methane. IPCC default assumption that
landfill gas is 50% methane)

16/12 = factor to convert from carbon to methane

R = Amount of landfill methane recovered. The value varies from year to year.

OX = 0.1 (10% of the methane that is not recovered is oxidized to CO2 in the upper
layers of the landfill cover.)

3.4.2 Wastewater treatment and sludge digestion

Finland uses IPCC’s default methodology to estimate methane emissions from wastewater
treatment (Technical Research Center of Finland 2001).

Emissions of CH4 = Organic load x B0 x MCF (Eq.13)

where: Organic Load is expressed (in Finland’s case) as COD for industrial
wastewaters and BOD in domestic wastewaters
B0 = the maximum methane generation potential – Finland uses the default
value given in the IPCC 1996 Revised Guidelines of 0.25 kg CH4/kg
COD or BOD. IPCC’s May 2000 Good Practices document revised
the default value to 0.25 kg CH4/kg COD and 0.6 kg CH4/kg BOD.
MCF = a weighted average value reflecting the methane generation potential
of the treatment plants in Finland relative to an anaerobic treatment
plant. For domestic wastewaters, Finland used an MCF of 0.025,
while for industrial wastewaters it used 0.005.

Emissions of N2O are estimated according to IPCC’s Revised 1996 Guidelines except that
Finland expands the scope to include nitrogen discharged not only in domestic sewage, but
also in industrial wastewaters and fish farm wastes (Technical Research Center of Finland
2001). The IPCC method estimates the N2O released from wastewater once it is discharged,
but assumes that N2O emissions from the treatment plants are negligible.

3.5 France – “Inventaire des émissions de gaz à effet de serre en France de 1990 à
1999” [Inventory of Greenhouse Gas Emissions in France for the Period 1990-
1999] (CITEPA 2000)

3.5.1 Landfills

France’s inventory document explains that the country has changed from IPCC’s all-in-one-
year approach (called the “zero-order” approach in the French document) to IPCC’s first
order decay model approach (CITEPA 2000, Section 4.6).

G - 17
Appendix 3 of the French Inventory document indicates that the estimates were developed
from national survey information describing the approximately 500 landfills in the country, a
number that is becoming smaller with time (CITEPA 2000). The data and modeling
assumptions were developed in consultation with France’s Agency for Environmental and
Energy Management (ADEME). The estimates were developed using the first order decay
model approach with emissions being a function of a number of factors including the
degradability of the waste, other waste characteristics such as the degree of compaction, and
the recovery of landfill gas (CITEPA 2000).

3.6 Japan - Information from the Japan Paper Association (JPA 2001)

3.6.1 Landfills

In preparing national inventories for greenhouse gases, the Ministry of Environment includes
methane and nitrous oxide from landfills containing “paper waste,” assumed to be sludges
and other process wastes from pulp and paper manufacturing (JPA 2001). The emission
factors are:

Methane = 151 kg/ton paper waste


Nitrous Oxide = 0.01 kg/ton paper waste

3.7 Sweden - Sweden’s National Inventory Report 2001 (Swedish Environmental


Protection Agency 2001)

3.7.1 Landfills

In “Sweden’s National Inventory Report 2001,” Sweden uses an approach very similar to
IPCC’s first order decay approach to estimate methane emissions from landfill. Waste
landfilled in 1952 and after is included in the analysis. The time factor in the rate equation is
adjusted slightly to correspond to an assumption that all waste is landfilled on July 1 of each
year. In addition, Sweden has developed country-specific values for a number of the
parameters used in the model (Swedish Environmental Protection Agency 2001).

The Swedish government has examined methane generation from pulp mill sludge landfills.
It uses a value of 45 kg CH4/ton of waste to represent the methane generation potential for
landfilled pulp mill sludge (Swedish Environmental Protection Agency 2001).

The other values used in the first order model for methane emissions from landfills are as
follows (Swedish Environmental Protection Agency 2001):
MCF before 1980 = 0.6
MCF in 1980 and after = 1.0
F (fraction of methane in landfill gas) = 0.5
DOCf (fraction of degradable organic carbon that is degraded to landfill gas) = 0.7
OX (fraction of non-collected gas that is oxidized in the surface layers of the landfill)
= 0.1

G - 18
t1/2 (“half-life of the methanogenesis”) = 7.5 years
k (first order rate constant assuming a half-life of 7.5 years) = 0.092/yr

3.7.2 Wastewater treatment and sludge digestion

The Swedish National Inventory Report indicates that “[e]missions of nitrous oxide from
wastewater handling are considered to be insignificant since almost all wastewater is treated
in aerobic conditions. Emissions from wastewater handling are therefore not reported”
(Swedish Environmental Protection Agency 2001).

Sludge-related emissions of methane from landfills are estimated as explained above. GHG
emissions from anaerobic sludge digestion are not discussed, although the report notes that
the landfill gas potential of already digested sludge is reduced by 50% (Swedish
Environmental Protection Agency 2001).

3.8 United Kingdom - UK Greenhouse Gas Inventory, 1990-1999 (AEA Technology


2001)

3.8.1 Landfills

The method used by the UK to develop national emission estimates for landfills is described
in Appendix 7 of the UK Greenhouse Gas Inventory (AEA Technology 2001).

“The UK method conforms to good practice since a first order decay (Tier 2)
methodology is used based on estimates of historical data on waste quantities,
composition and disposal practices over several decades.”

“The UK method is based on equations 4 and 5 in the Revised 1996 IPCC


guidelines... which are compatible with equations 5.1 and 5.2 in the Good Practice
Guidance.... A slightly different version of equation 5.1 is used, which takes into
account the fact that the model uses a finite time interval (one year)…”

“The UK method divides the waste stream into four categories of waste: rapidly
degrading, moderately degrading, slowly degrading and inert waste. These categories
each have a separate decay rate. The decay rates are based on data from the
Netherlands and range from 0.05 (slowly degrading waste) to 0.185 (rapidly
degrading waste), which lie within the range of 0.03 to 0.2 quoted in the Good
Practice Guidance.”

“The model extends back to 1945, which gives a time period of around 4 half lives
for the slowest of the three decay rates (0.05, half life 14 years). This lies within the
range of 3 to 5 half lives recommended by the Good Practice Guidance.”

“The model distinguishes between four separate categories of landfill site with
different degrees of gas collection control:
· closed sites;

G - 19
· sites with no gas collection;
· sites with limited collection;
· sites with comprehensive collection.”

“Each site type has different gas collection and oxidation rates. As recommended, the
model attempts to take into account changes in landfill practice over past decades by
altering the proportion of waste disposed of to each of these categories of site in past
years, and also by modifying the gas collection rate over time where appropriate. The
model also simulates retrofitting of sites, i.e., upgrading from a site with no gas
control or limited gas control to one with comprehensive gas control.”

“The estimates of historical waste disposal and composition data are based on various
data sources… As recommended in the Good Practice Guidance, estimates for
municipal waste are based on population where data is absent.”

“As recommended in the Good Practice Guidance, the estimates of waste disposal
quantities include commercial and industrial waste, demolition and construction
waste and sewage sludge, as well as municipal waste. For industrial and commercial
waste, the data are based on national estimates from a recent survey, although the
survey was incomplete at the time of finalizing the model estimates. The data were
extrapolated to cover past years based on employment rates in the industries
concerned.”

“All sites in the UK are managed, and therefore have a methane correction factor of
1.0. However, as described above, differences in oxidation rates have been simulated
by the practice of dividing waste disposal sites into four types as described above.”

“Degradable organic carbon (DOC) was estimated based on a national study, as


recommended in the Good Practice Guidance. However the figures used were based
on expert opinion rather than measured data.”

“The fraction of degradable organic carbon dissimilated (DOCF) was assumed to be


60%. At the time when the model was set up, the IPCC recommended default value
was 77%, but there were indications that this could be an overestimate, so a lower
figure was used. The new IPCC recommended range quoted in the Guidance is 50-
60%.”

“The fraction of CH4 in landfill gas is generally taken to be 50%, which is in line with
the Guidance. For old shallow sites it is taken to be 30% to reflect a higher degree of
oxidation. The fraction of methane recovered was assumed to be 85% for sites with
full gas control and 40% for sites with limited gas control. The estimates are not
derived from metering data, as recommended by the Guidance, as such data were not
readily available at the time of the study. A panel of UK industry experts selected the
figures.”

G - 20
“The oxidation factor is assumed to be 10% for all site types. Recovered methane is
subtracted before applying the oxidation factor. This is in line with the Guidance….”

“Neither the GHGI nor the NAEI reports carbon dioxide emissions from the
anaerobic decay of landfilled waste since this is considered to be part of the carbon
cycle and is not a net source. The estimates include the contribution of sewage
disposed of to landfill.”

3.8.2 Wastewater treatment

The method used by the UK to develop national estimates of GHG emissions from
wastewater treatment is described in Appendix 7 of the UK Greenhouse Gas Inventory (AEA
Technology 2001). That document explains that the UK does not attempt to estimate
emissions from private wastewater treatment plants. The method used to estimate emissions
from publicly owned treatment plants is described as follows (AEA Technology 2001).

“The methodology of the UK model differs in some respects from the IPCC default
methodology. The main differences are that it considers wastewater and sewage
together rather than separately. It also considers domestic, commercial and industrial
wastewater together rather than separately. Emissions are based on empirical
emission factors derived from the literature expressed in kg CH4/tonne dry solids
rather than the BOD default factors used by IPCC. The model however complies with
the IPCC Good Practice Guidance as a national model....”

“The basic activity data are the throughput of sewage sludge through the public
system. The estimates are based on the UK population connected to the public sewers
and estimates of the amount of sewage per head generated. From 1995 onwards the
per capita production is a projection.... The main source of sewage activity data is the
UK Sewage Survey.... Emissions are calculated by disaggregating the throughput of
sewage into 14 different routes. The routes consist of different treatment processes
each with its own emission factor. The treatment routes and emission factors are
shown in [the table below]…”

“The model accounts for recovery of methane and its subsequent utilization and
flaring by estimating the proportion of anaerobic digester emissions that are
recovered.”

“Nitrous oxide emissions from the treatment of human sewage are based on the
IPCC... default methodology.”

G - 21
Table 3. Methane Emission Factors for Sludge Handling kg CH4/Mg dry solids,
(AEA Technology 2001, from Table 1 of Appendix 7)

Sludge Handling System Gravity Long-term Anaerobic Agricultural


1
Thickening Storage Digestion2 Land
Anaerobic digestion to agriculture 0.72 143 5
Digestion, drying, agriculture 0.72 143 5
Raw sludge, dried to agriculture 0.72 20
Raw sludge, long term storage (3m), 0.72 36 20
agriculture
Raw sludge, dewatered to cake, 0.72 20
agriculture
Digestion, to incinerator 0.72 143
Raw sludge, to incinerator 0.72
Digestion , to landfill 0.72 143
Compost, to agriculture 0.72 5
Lime raw sludge, to agriculture 0.72 20
Raw Sludge , to landfill 0.72
Digestion , to sea disposal 0.72 143
Raw sludge to sea disposal 0.72
Digestion to beneficial use(e.g., land 0.72 143 5
reclamation)
1
An emission factor of 1 kg/t is used for gravity thickening, Around 72% of sludge is
gravity thickened hence an aggregate factor of 0.72 kg CH4/Mg is used.
2
The factor refers to methane production, however it is assumed that 121.5 kg CH4/Mg
is recovered or flared.

3.9 United States - U.S. EPA Inventory of U.S. Greenhouse Gas Emissions and
Sinks: 1990-1999 (USEPA 2001a) and Anthropogenic Methane Emissions in the
United States: Estimates for 1990, Report to Congress (USEPA 1993)

3.9.1 Landfills

EPA’s analysis of GHG emissions from landfills is focused on methane from municipal solid
waste landfills, although estimates are also given for industrial landfill emissions of methane.
The following description of the methodology is taken from EPA’s “Inventory of U.S.
Greenhouse Gas Emissions and Sinks: 1990-1999” (USEPA 2001a, page 7-4).

“Based on available information, methane emissions from landfills


were estimated to equal the methane produced from municipal
landfills, minus the methane recovered and combusted, minus the

G - 22
methane oxidized before being released into the atmosphere, plus the
methane produced by industrial landfills.”

“The methodology for estimating CH4 emissions from municipal


landfills is based on a model that updates the population of U.S.
landfills each year. This model is based on the pattern of actual waste
disposal by each individual landfill surveyed by the EPA’s Office of
Solid Waste in 1987. A second model was employed to estimate
emissions from the landfill population. [This model is described in
USEPA 1993.] For each landfill in the data set, the amount of waste-
in-place contributing to methane generation was estimated using its
year of opening, its waste acceptance rate, year of closure, and design
capacity. Data on national waste disposed in landfills each year was
apportioned by landfill. Emissions from municipal landfills were then
estimated by multiplying the quantity of waste contributing to
emissions by emission factors…. “

“The estimated landfill gas recovered per year was based on updated
data collected from vendors of flaring equipment, and a database
compiled by the EPA's Landfill Methane Outreach Program
(LMOP)….”

“Emissions from industrial landfills were assumed to be equal to 7


percent of the total methane emissions from municipal landfills. The
amount of methane oxidized was assumed to be 10 percent of the
methane generated. To calculate net methane emissions, both methane
recovered and methane oxidized were subtracted from methane
generated at municipal and industrial landfills.”

The “model” described in “Anthropogenic Methane Emissions in the United States:


Estimates for 1990, Report to Congress,” is a two parameter model for MSW landfills
(USEPA 1993). It is based on an empirical analysis of gas generation data from more than
85 US MSW landfills and estimates methane generation based on the mass of waste in place
and the ambient rainfall.

EPA also has a model, however, that is equivalent to the first order decay model suggested
by IPCC. EPA’s first order model is described in its compilation of emission factors, AP-42
(USEPA 1998). EPA calls its model the Landfill Air Emissions Estimation Model
(LAEEM). NCASI has reviewed EPA’s normal approach for estimating methane emissions
from landfills and the results are reported in NCASI Technical Bulletin No. 790, September
1999 (NCASI 1999). The following material is taken mostly from that source.

LAEEM is a PC-based automated estimation tool, operating in a Windows™ environment,


for calculating uncontrolled air emissions from municipal solid waste (MSW) landfills,
available from the EPA Office of Air Quality Planning and Standards.

G - 23
LAEEM incorporates the “Scholl Canyon” model, which is a first order, single stage model
identical to IPCC’s. Kinetic rate coefficients were empirically adjusted to reflect changes in
refuse moisture content and other landfill conditions. The Scholl Canyon model assumes that
the gas production rate is at its peak upon initial waste placement and that anaerobic
conditions are established immediately. Gas production is then assumed to decrease
exponentially as a first order decay. The model allows for division of the landfill into
modules (annual refuse accumulations) to account for different ages of the refuse
accumulated over time.

A default first order rate constant for methane generation (“k”) of 0.04/yr is recommended
for areas receiving 25 inches or rain or more per year while a value of 0.02/yr is
recommended for drier areas. A default value for methane generation potential, L0, of 100
m3 methane/Mg waste is recommended (USEPA 1998). An examination of the source of
these recommendations by NCASI led to the conclusion that EPA’s default values (derived
from studies of MSW landfills) is probably too high for pulp and paper industry landfills
(NCASI 1999).

NCASI is generating industry-specific values for L0 and k in both the laboratory and the
field. Studies to determine values for L0 are underway in the laboratory for five paper
industry sludges (including mixing with causticizing wastes and fly ash) and are nearing
completion. Because the rate constant is sensitive to field conditions such as temperature and
moisture, NCASI has initiated field experiments to investigate the rate constant that can be
expected under real world conditions. Results of these experiments for rate constants will be
available in two to four years. Preliminary examination of the data generated in the studies
supports the conclusions drawn in Technical Bulletin No. 790. (NCASI 1999)

3.9.2 Wastewater treatment and anaerobic sludge digestion

EPA indicates that it used the IPCC methodology for estimating methane emissions from
wastewater treatment. This involved estimating the amount of wastewater organic matter
produced and multiplying that amount by an emission factor. EPA estimated methane
emissions only from anaerobic treatment operations, which were assumed to be handling
15% of the domestic wastewater BOD generated in the US (USEPA 2001a).

The emission factor used by EPA is 0.6 kg CH4/kg BOD5. EPA cites IPCC’s May 2000
“Good Practice Guidance and Uncertainty Management in National Greenhouse Gas
Inventories” as the source of the 0.6 kg CH4/kg BOD emission factor.

EPA also develops separate estimates of methane emissions from pulp and paper industry
wastewater treatment operations. The estimate was developed by assuming that one percent
of the pulp and paper mill wastewater treated in mill-owned treatment plants is treated
anaerobically in systems that allow the methane to escape to the atmosphere. The emission
factor used was the same one EPA used for municipal wastewater treatment, 0.6 kg CH4/kg
BOD5. For pulp and paper mills, however, EPA estimated methane emissions by multiplying
the emission factor by the amounts of BOD removed in treatment. The IPCC factor of 0.6 kg
CH4/kg BOD is based, however, on organic matter in untreated wastewater. Because of the

G - 24
high removal efficiencies for BOD in wastewater treatment, the discrepancy is relatively
unimportant (USEPA 2001).

3.10 WRI/WBCSD – The Greenhouse Gas Protocol, October 2001 (WRI 2001a)

The WRI/WBCSD Protocol divides GHG emissions from waste management into those from
company-owned sources (included in Scope 1 of the WRI/WBCSD Protocol) and those from
sources owned by other entities (included in Scope 3). The tools for estimating these
emissions are still under development (WRI 2001a).

G - 25
ANNEX H

ALLOCATING GREENHOUSE GAS EMISSIONS FROM PARTLY OWNED OR


PARTLY CONTROLLED SOURCES – OVERVIEW OF METHODS IN EXISTING
PROTOCOLS

1.0 INTRODUCTION

Those protocols that are focused on geographic or national inventories have no need to
address the question of corporate “ownership” of emissions from jointly-owned or –operated
facilities. This leaves only a few of the protocols that address the question. The
WRI/WBCSD GHG Protocol offers the most comprehensive discussion of this issue and
should be viewed as a primary resource for companies having to determine the ownership of
emissions from sources that are partly owned or controlled.

1.1 Canada - FPAC’s 2001 Greenhouse Gas Action Plan Guidelines

FPAC’s 2001 Greenhouse Gas Action Plan Guidelines do not directly address the question of
how to allocate emissions from jointly owned or operated facilities except for combined heat
and power projects reported under Canada’s Baseline Protection Initiative. In this case,
FPAC indicates that “[p]arties involved in co-generation projects may decide between
themselves how to share direct reductions” (FPAC 2001).

1.2 United States - U.S. Department of Energy 1605(b) program guidelines

This voluntary project-level reporting program does not specify how to allocate emissions or
emissions reductions involving more than one corporate entity. The Department of Energy
leaves it to the participating companies to decide this, requiring only that the allocation be
transparent (USDOE 2000).

1.3 WRI/WBCSD - The Greenhouse Gas (GHG) Protocol Initiative

In October 2001, WRI/WBCSD released “The Greenhouse Gas Protocol – a corporate


accounting and reporting standard.” The following material is taken entirely from that
document, which companies facing this issue will want to obtain and read in its entirety
(WRI 2001a).

“When accounting for GHG emissions from partially-owned entities/facilities,


it is important to draw clear organizational boundaries, which should be
consistent with the organizational boundaries which have been drawn up for
financial reporting purposes. “

“Financial reporting is based upon the concepts of ‘control’ and ‘influence’.


The concepts of ‘control’ and ‘influence’ are often defined and applied
differently according to a company’s specific financial accounting and

H-1
reporting policies/practices. Where possible, it makes sense to follow
company-specific distinctions already in place for financial accounting,
provided these are explicitly explained and followed consistently. When
applying these concepts the underlying assumption of ‘substance over form’
should be followed. This assumption is based on the premise that GHG
emissions should be accounted and reported in accordance with the
company’s substance and economic reality and not merely its legal form.”

“For the purpose of applying the concepts of ‘control’ and ‘significant


influence’ to GHG accounting, the following definitions may prove helpful.”

“Control is defined as the ability of a company to direct the operating policies


of another entity/facility. Usually, if the company owns more than 50 percent
of the voting interests, this implies control. The holder of the operating license
often exerts control, however, holding the operating license is not a sufficient
criteria for being able to direct the operating policies of an entity/facility. In
practice, the actual exercise of dominant influence itself is enough to satisfy
the definition of control without requiring any formal power or ability through
which it arises.”

“Significant influence: the issue of whether a company has significant


influence over an entity/facility is likely to have been already established by
the company-specific financial accounting and reporting policies/practices.
However, where it is necessary to determine if one company exerts significant
influence over an entity/facility, the following factors should be considered:
• the company owns voting interests of between 20 and 50 percent
• the company has the power to participate in the entity’s/facility’s financial
and operating policy decisions
• the company has a long-term interest in the entity/facility”

“...GHG emissions from entities/facilities that are not under significant


influence or control (e.g., the company owns less than 20 percent of the voting
interests) are generally not reported. This is consistent with financial
accounting standards where a company would only recognize revenue if
dividends were paid or a loss incurred if the asset was impaired. However, it is
recognized that GHG emissions are different in their nature and it may be
appropriate for these to be reported to properly reflect the company’s overall
GHG emissions. If this is the case it is important to state this in the public
report.”

“Companies should preferably account for and report their GHG emissions
according to the framework presented in [the following table]. This
framework is set out to provide GHG emissions information in a transparent
manner on the basis of control/influence and equity share basis. Equity share
is defined as the percentage of economic interest in/benefit derived from an
operation. This approach increases the usability of GHG information for

H-2
different users and aims, as far as possible, to mirror the approach adopted by
financial accounting and reporting standards.”

“Where there is a contractual arrangement that covers GHG emissions, the


company should defer to this for the purposes of emissions allocation.”

“…What constitutes ‘control’ and ‘significant influence’ may not always be


obvious. The definitions provided by different financial accounting and
reporting standards such as the International Accounting Standards (IAS) and
US Generally Accepted Accounting Principles (US GAAP) do not always
converge. Therefore, when accounting for GHG emissions from partially-
owned entities/facilities you should follow, as closely as possible, the
distinctions of ‘control’ and ‘significant influence’ as applied by your
company for the financial consolidation of such entities/facilities.”

“…GHG emissions reduction initiatives, and regulatory and trading schemes


often focus on control rather than ownership. The company holding the
operating license might be asked to report 100 percent of the operations’ GHG
emissions. It is therefore important to distinguish between operations
controlled on the basis of the operating license and those controlled on the
basis of majority voting interest or other reasons.”

The WRI/WBCSD Protocol recommends that were possible, companies report emissions
using both the equity share and control/influence methods to allocate emissions from
partially owned sources. The allocation recommendations are summarized in the following
table (WRI 2001a). The WRI/WBCSD report contains several examples of how to allocate
emissions based on control/influence and equity share.

H-3
Table 3. Accounting for GHG Emissions on the Basis of Control and Equity Share
(Table 1 from WRI 2001a)

Category What to Report


Reporting for Control
Controlled entities/facilities Wholly owned 100% of GHG emissions
Emissions from those entities/facilities, which are
defined as being controlled. It is likely that this will
Not wholly owned 100% of GHG emissions
already be determined by company-specific
but controlled
financial accounting policies/practices.
This category includes entities/facilities that are:
Jointly controlled Equity share of GHG
· wholly owned
emissions
· not wholly owned, but controlled
· jointly controlled assets/entities
The concept of jointly controlled assets/entities will
have to be considered based on the specific business
and industry context.
Reporting for Equity Share
A. Controlled entities/facilities Equity share of GHG emissions
Emissions from those entities/facilities,
If there is a specific contractual arrangement that
which are defined as being controlled. It is
covers the division of earnings/production, that
likely that this will already be determined
arrangement should be considered. This is likely
by company-specific financial accounting
to be most prominent in the upstream oil and gas
policies/practices.
industry, and is determined by company-specific
This category includes entities/facilities financial accounting policies/practices.
that are:
• wholly owned
• not wholly owned, but controlled
• jointly controlled assets/entities
B. Significant influence – associated Equity share of GHG emissions
entities/facilities
Emissions from entities/facilities over
If there is a specific contractual arrangement that
which the reporting company has
covers the division of earnings/production, that
significant influence but does not control.
arrangement should be considered. This is likely
It is likely that this will already be
to be most prominent in the upstream oil and gas
determined by company specific financial
industry, and is determined by company-specific
accounting policies/practices.
financial accounting policies/practices.
Equity share of GHG emissions Equity share of GHG emissions from
entities/facilities that are controlled or under
significant influence ( A + B )

H-4
ANNEX I

SUPPORTING INFORMATION

1.0 ESTIMATES OF CLIMATE-NEUTRAL CO2 EMISSIONS FROM BIOMASS


COMBUSTION

Pulp and paper mills generate approximately two thirds of their energy needs from
greenhouse gas-neutral biomass fuels recovered from the industry’s waste and process
streams. Energy-rich biomass – derived from wood chips, bark, sawdust and pulping liquors
recovered from the harvesting and manufacturing processes – is atmospheric carbon dioxide
sequestered by trees during growth and transformed into organic carbon substances. When
these biomass fuels are burned, the CO2 emitted during the manufacturing and combustion
processes is the atmospheric carbon dioxide that was sequestered during growth of the tree;
hence, there is no net contribution to the atmospheric CO2 level. This carbon cycle is a
closed-loop. New tree growth keeps absorbing atmospheric carbon dioxide and maintains
the cycle.

Any increases or decreases in the amount of carbon sequestered by the forests are accounted
for in the comprehensive forest accounting system. This is the approach generally prescribed
for national inventories by the United Nations Framework Convention on Climate Change.
Most international protocols including that of the Intergovernmental Panel on Climate
Change (IPCC) have adopted the convention set out by the United Nations. The IPCC has
stated that emissions from biomass do not add to atmospheric concentrations of carbon
dioxide (IPCC 1997a,c).

1.1 Estimated Biomass Emissions

The information reported below on biomass emissions is being supplied:


· To ensure that readers understand the entity’s overall energy profile in terms of both
greenhouse gas emissions and non-greenhouse gas emissions, and
· To provide awareness and understanding of how greenhouse-gas-neutral biomass fuels
are generated and used in the manufacture of pulp and paper.

Tables 1 through 3 on the following pages can be used to estimate emissions of climate-
neutral CO2 emissions from biomass combustion and Table 4 can be used to record the
results.

Also, note that the information reported below is in conformance with the general greenhouse
gas protocol designed by the World Resources Institute and the World Business Council for
Sustainable Development (WRI 2001a). Users of these calculation tools may elect to modify
the format and type of information presented based on specific facility or company needs.

I-1
Table 1. Estimating Emissions of Climate-neutral CO2 from Combustion of Wood, Bark, and other Biomass Fuels (except pulping
liquors which are addressed in Table 2 below)
Step 1 Step 2 Step 3
A B C D E F
Quantity of Unit used to CO2 emission Unit of CO2 CO2 CO2 emissions in
fuel burned measure quantity factor: emission factor emissions in metric tons Carbon/yr
of fuel use [default value is: kg CO2/yr
[Note: Be solid biomass:
careful not to 109*
mix HHVs and kg CO2/GJ LHV]
LHVs.]

Source Description Fuel Type E=A*C F = E*12/44 / 1000


Example: Bark Boiler Bark 500,000 GJ (LHV) 109 kg CO2 / GJ LHV 54,500,000 14,900

Biomass Carbon Released as climate-neutral CO2 from Combustion of Wood , Bark, or other biomass
(except pulping liquors which are addressed in Table 2 below)
Amount of climate-neutral biomass-derived CO2 included in above number that is exported
(e.g., to PCC plant) rather than being emitted - Optional information
* Emission factor for solid biomass from Intergovernmental Panel on Climate Change (IPCC). 1997. Revised 1996 IPCC Guidelines for National Greenhouse
Gas Inventories: Reference Manual (Volume 3). Table 1.1, corrected for 1% unburned carbon (USEPA 2001b)

NOTE: All listed emission factors are based on LHV – LHV assumed to be 95% of HHV for biomass fuels
Table 2. Estimating Emissions of Climate-neutral CO2 from Pulping Liquors and Kraft mill lime recovery
Note that because the emission factors for spent pulping liquors are based on the carbon content of the liquors (assuming a one-percent correction for unoxidized
carbon), the liquor emission factor will include any carbon exiting with smelt from a recovery furnace. Therefore, for kraft mills, the liquor emission factors
estimate biomass carbon emissions from both the recovery furnace and from the lime kiln.
Companies have the option of estimating the amounts of climate-neutral biomass-derived CO2 that is exported rather than being returned to the atmosphere. If the
exports are from a lime kiln or calciner where fossil fuels are being used, as a general approximation the exports of climate-neutral biomass-derived CO2 will be
twice the exports of fossil fuel-derived CO2 (see "CO2 Imports and Exports" worksheet). Where exports consist of gases from a lime kiln or calciner that is
burning only biomass fuels, all of the exports are carbon-neutral biomass-derived CO2.
Step 1 Step 2 Step 3
A B C D E F
Quantity of Unit used to CO2 emission Unit of CO2 CO2 CO2 emissions
fuel burned measure quantity of factors: emission factor emissions in in metric tons
fuel use [default values kg CO2/yr Carbon/yr
[Note: Be careful are listed in Table
not to mix HHVs 3 below;
and LHVs.] kg CO2/GJ LHV]
Source Description Fuel Type E=A*C F = E*12/44 / 1000
Example: Recovery Kraft Pulping 100,000 GJ (LHV) 89.9 kg CO2 / GJ LHV 8,990,000 2,450
Furnace Liquor, North
American SW

Biomass Carbon in Pulping Liquors Released as climate-neutral CO2 from the recovery furnace and lime kiln or calciner
Amount of climate-neutral biomass-derived CO2 included in above number that is exported
(e.g., to PCC plant) rather than being emitted - Optional information

NOTE: All listed emission factors are based on LHV – LHV assumed to be 95% of HHV for biomass fuels
Table 3. Suggested pulping liquor default emission factors for climate-neutral CO2
- includes emissions from both the recovery furnace and lime kiln/calciner -

Typical Carbon Typical Energy Content – Calculated Energy Climate-neutral CO2


Content HHV Content - LHV Emission Factor
Type of Pulping Liquor Wood Furnish [percent, dry basis] [GJ HHV / metric ton dry [GJ LHV / metric ton [kg CO2 / GJ LHV]
solids] dry solids]
Kraft black liquor* Scandinavian Softwood 35 14.2 13.5 89.9
Kraft black liquor* Scandinavian Hardwood 32.5 13.5 12.8 87.8
Kraft black liquor* North American Softwood 35 14.2 13.5 89.9
Kraft black liquor* North American Hardwood 34 13.9 13.2 89.2
Kraft black liquor* Tropical Eucalyptus 34.8
Kraft black liquor* Tropical Mixed Woods 35.2 14.1 13.4 91.1
Kraft black liquor* Bagasse 36.9 14.8 14.1 90.9
Kraft black liquor* Bamboo 34.5 14.1 13.4 89.2
Kraft black liquor* Straw 36.5 14.7 14.0 90.6
Semi-Chemical to be determined
Sulfite to be determined

* kraft black liquor default emission factors are from: Chapter 1- Chemical Recovery, by Esa Vakkilainen. 1999. In: Papermaking Science and
Technology, Book 6B: Chemical Pulping. Gullichsen, J., and Paulapuro, H. (eds.). Helsinki, Finland: Fapet Oy
Table 4. Direct Emissions of Climate-neutral CO2 from Biomass

Write “NA” to show an item is not applicable.


Where emissions have been determined to be insignificant or non-material, write Direct Emissions of
“NM” and explain the basis for the determination in a footnote. Climate-neutral CO2
– metric tons -

1 Biomass– fueled boilers (from Table 1 in Annex I)

2 Pulping liquor-derived CO2 (from Table 2 in Annex I)

3 Total Direct Emissions of Climate-neutral CO2 (Sum of line 1 and 2)


Explain the method used to determine ownership of emissions from sources not completely owned by the
company. Use the WRI/WBCSD GHG Protocol for guidance on determining ownership.

Include any other information that is needed to understand the inventory results:

I-5
ANNEX J

REFERENCES FOR ANNEXES A THROUGH H

AEA Technology. 2001. UK Greenhouse Gas Inventory 1990-1999: Annual Report for
submission under the Framework Convention on Climate Change. National Environmental
Technology Centre. Oxfordshire, England.
http://ariadne.aeat.co.uk/netcen/airqual/reports/ghg/ghg2.html (26 Nov. 2001)

Australian Greenhouse Office (AGO). 1998. Energy Workbook for Fuel Combustion
Activities (Stationary Sources) – Workbook 1.1 with Supplements - 1998. Australian
Greenhouse Office. Canberra, Australia;
http://www.greenhouse.gov.au/inventory/methodology/98workbook1.pdf (26 Nov. 2001)

Australian Greenhouse Office (AGO). 1999. Workbook for Fuel Combustion Activities
(Stationary Sources) – Workbook 1.1 – 1999 Supplement. Australian Greenhouse Office.
Canberra, Australia.
http://www.greenhouse.gov.au/inventory/inventory/pdfs/methodstat99.pdf (26 Nov. 2001)

Australian Greenhouse Office (AGO). 2001a. National Greenhouse Gas Inventory – 1999.
Australian Greenhouse Office. Canberra, Australia.
http://www.greenhouse.gov.au/inventory/inventory/latest_inventory.html (26 Nov. 2001)

Australian Greenhouse Office (AGO). 2001b. Greenhouse Challenge – Factor and Methods
Workbook, Version 3 – December 2001. Australian Greenhouse Office, Canberra, Australia,
http://www.greenhouse.gov.au/challenge/html/member-tools/factorsmethod.html (15 Feb.
2002)

Australian Paper Industry Council (APIC). 2002. Environmental Performance Indicators.


http://www.apic.asn.au/environment/default.htm (15 Feb. 2002)

Axelsson, L. 2001. Personal communication between Lars-Erik Axelsson of the Swedish


Forest Industries Federation and Reid Miner, NCASI. September 2001

Barnard, G. 2001. Personal communication between Graham Barnard of the Paper


Federation of Great Britain and Reid Miner of NCASI. September 2001

Canada’s Climate Change Voluntary Challenge and Registry Inc. (VCR). 1999. Registration
Guide 1999 – Canada’s Climate Change Voluntary Challenge and Registry Inc.
http://www.vcr-mvr.ca/Downloads/PDF/VCR-RegistrationGuide_e.pdf (26 Nov. 2001)

Centre Interprofessionnel Technique d’Etudes de la Pollution Atmosphérique (CITEPA).


2000. Inventaire des émissions del gaz à effet de serre en France de 1990 à 1999. Centre
Interprofessionnel Technique d’Etudes de la Pollution Atmosphérique. Paris, France.
http://www.citepa.org/emissions/france_objectifs/Emissions_France_Ipcc_2000.pdf (26 Nov.
2001)

J-1
Chilean National Environmental Commission. 1999. First National Communication to the
Conference of the Parties to the United Nations Framework Convention on Climate Change.
Chilean National Environmental Commission. Santiago, Chile.
http://www.unfccc.de/resource/docs/natc/chinc1.pdf (26 Nov. 2001)

Environment Canada. 1999. Canada’s Greenhouse Gas Inventory – 1997 Emissions and
Removals with Trends. Environment Canada. Ottawa, Canada.
http://www.ec.gc.ca/pdb/ghg/ghg_docs/gh_eng.pdf (26 Nov. 2001)

European Environment Agency (EEA). 1999. EMEP/CORINAIR Atmospheric emission


inventory guidebook - Second edition. Copenhagen, Denmark.
http://reports.eea.eu.int/EMEPCORINAIR/en (26 Nov. 2001)

Finnish Forest Industries Federation. 2001. Environmental Report – Statistics for 2000.
www.forestindustries.fi/index_2.htm (26 Nov. 2001)

Forest Products Association of Canada (FPAC). 2001. 2001 Greenhouse Gas Action Plan
Guidelines. Forest Products Association of Canada. Montreal, Canada

Fortum Power and Heat Oy (Fortum). 2001. Methane and Nitrous Oxide Emissions in the
Finnish Energy Production. Fortum Power and Heat Oy. Fortum, Filand, Technology

Hough, G.. 1985. Chemical Recovery in the Alkaline Pulping Processes. TAPPI Press.
Atlanta GA

International Energy Agency (IEA). 1998. The Role of Bioenergy in Greenhouse Gas
Mitigation. International Energy Agency. Paris.
http://www.ieabioenergy.com/media/6_pospapa4.pdf (14 Feb. 2002)

Intergovernmental Panel on Climate Change (IPCC). 1997a. Revised 1996 IPCC Guidelines
for National Greenhouse Gas Inventories: Reporting Instructions (Volume 1). IPCC National
Greenhouse Gas Inventory Program. http://www.ipcc-nggip.iges.or.jp/public/gl/invs4.htm
(26 Nov. 2001)

Intergovernmental Panel on Climate Change (IPCC). 1997b. Revised 1996 IPCC Guidelines
for National Greenhouse Gas Inventories: Workbook (Volume 2). IPCC National Greenhouse
Gas Inventory Program. http://www.ipcc-nggip.iges.or.jp/public/gl/invs5.htm (26 Nov. 2001)

Intergovernmental Panel on Climate Change (IPCC). 1997c. Revised 1996 IPCC Guidelines
for National Greenhouse Gas Inventories: Reference Manual (Volume 3). IPCC National
Greenhouse Gas Inventory Program. http://www.ipcc-nggip.iges.or.jp/public/gl/invs6.htm
(26 Nov. 2001)

J-2
Intergovernmental Panel on Climate Change (IPCC). 2000. Good Practice Guidance and
Uncertainty Management in National Greenhouse Gas Inventories. IPCC National
Greenhouse Gas Inventory Program.
http://www.ipcc-nggip.iges.or.jp/public/gp/gpgaum.htm (26 Nov. 2001)

Japan Paper Association (JPA). 2001. Various personal communications between Masao
Taniguchi of JPA and Reid Miner of NCASI. Japan Paper Association. Tokyo Japan

Karessuo, A. 2001. Personal Communication between Anu Karessuo of the Finnish Forest
Industries Federation and Reid Miner of NCASI. September 2001

Miner, R. and Upton, B.. 2002. Methods for estimating greenhouse gas emissions from lime
kilns at kraft pulp mills. Energy, Volume 27 No. 8, pp. 729-738 (Sept. 2002) Elsevier
Science, New York, NY

National Council for Air and Stream Improvement (NCASI). 1980. A Study of Wood-Residue
Fired Power Boiler Total Gaseous Non-Methane Organic Emissions in the Pacific Northwest
- Air Quality Technical Bulletin No.109. NCASI. Research Triangle Park, NC

National Council for Air and Stream Improvement (NCASI). 1981. A Study of the Kraft
Process Lime Kiln Total Gaseous Non-Methane Organic Emissions - Technical Bulletin No.
358. NCASI. Research Triangle Park, NC

National Council for Air and Stream Improvement (NCASI). 1985. Volatile Organic Carbon
Emissions from Wood Residue Fired Power Boilers in the Southeast - Technical Bulletin
No.455. NCASI. Research Triangle Park, NC

National Council for Air and Stream Improvement (NCASI). 1999. Evaluation of the EPA-
Recommended Approach to Predicting Air Emissions from Pulp and Paper Landfills -
Technical Bulletin No.790. NCASI. Research Triangle Park, NC

Pulp and Paper Manufacturers Federation of Australia (PPMFA). 2001. Australian Pulp and
Paper Industry Environmental Performance Indicators Survey Form. Pulp and Paper
Manufacturers Federation of Australia. Manuka. Australia

Swedish Environmental Protection Agency. 2001. Sweden’s National Inventory Report 2001
– Submitted under the United National Convention on Climate Change. Swedish
Environmental Protection Agency. Stockholm, Sweden.
http://www.environ.se/dokument/fororen/klimat/klimat/fcccdata/NIR2001.pdf (26 Nov.
2001)

Technical Research Center of Finland. 2001. Greenhouse Gas Emissions and Removals in
Finland. Technical Research Center of Finland. Espoo, Finland.
http://www.vyh.fi/eng/environ/state/air/emis/ghg/methodol.htm (26 Nov. 2001)

J-3
United States Department of Energy (USDOE). 2000. General Guidelines and Supporting
Documents Establishing the Voluntary Reporting of Greenhouse Gases Program. US
Department of Energy. Energy Information Administration. Washington D.C.
http://www.eia.doe.gov/oiaf/1605/guidelns.html (26 Nov. 2001)

United States Department of Energy (USDOE). 2001a. Updated State-level Greenhouse Gas
Emission Factors for Electricity Generation. US Department of Energy. Energy Information
Administration. Washington, D.C. ftp://ftp.eia.doe.gov/pub/oiaf/1605/cdrom/pdf/e-
supdoc.pdf (26 Nov. 2001)

United States Department of Energy (USDOE). 2001b. Average Electricity Emission Factors
by State. US Department of Energy. Energy Information Administration. Washington, D.C.
http://www.eia.doe.gov/oiaf/1605/e-factor.html (26 Nov. 2001)

United States Environmental Protection Agency (USEPA). 1993. Anthropogenic Methane


Emissions in the United States: Estimates for 1990, Report to Congress. EPA 430-R-93-003.
United States Environmental Protection Agency. Washington, D.C.

United States Environmental Protection Agency (USEPA). 1998. AP-42 Emission Factors
for Municipal Solid Waste Landfills – Supplement E, November 1998. United States
Environmental Protection Agency. Washington, D.C.
http://www.epa.gov/ttn/chief/ap42/ch02/final/c02s04.pdf (26 Nov. 2001)

United States Environmental Protection Agency (USEPA). 2001a. Inventory of U.S.


Greenhouse Gas Emissions and Sinks: 1990-1999. EPA 236-R-01-001. United States
Environmental Protection Agency. Washington, D.C.
http://www.epa.gov/globalwarming/publications/emissions/us2001/index.html (26 Nov.
2001)

United States Environmental Protection Agency (USEPA). 2001b. AP-42 Emission Factors
for Wood Residue Combustion in Boilers – Supplement G, July 2001. United States
Environmental Protection Agency. Washington, D.C.
http://www.epa.gov/ttn/chief/ap42/ch01/final/c01s06.pdf (26 Nov. 2001)

World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001a. The Greenhouse Gas Protocol: a corporate accounting and reporting standard.
World Resources Institute. Washington, D.C. http://www.ghgprotocol.org/standard/ghg.pdf
(26 Nov. 2001)

World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001b. Calculating CO2 emissions from the combustion of standard fuels and from
electricity/steam purchase - Guide to calculation worksheets. World Resources Institute.
Washington, D.C. http://www.ghgprotocol.org/standard/stationarycombustion.doc (26 Nov.
2001)

J-4
World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001c. Calculating CO2 emissions from the combustion of standard fuels and from
electricity/steam purchase - Calculation worksheets. World Resources Institute. Washington,
D.C. http://www.ghgprotocol.org/standard/stationarycombustion.xls (26 Nov. 2001)

World Resources Institute (WRI) and World Business Council for Sustainable Development.
2001d. Calculating CO2 emissions from Mobile Sources -Guide to calculation worksheets.
World Resources Institute. Washington, D.C.
http://www.ghgprotocol.org/standard/mobile.doc (26 Nov. 2001)

J-5

Anda mungkin juga menyukai